Variational Problems in Differential Geometry
Variational Problems in Differential Geometry
The titles below are available from booksellers, or from Cambridge University Press at http://www.cambridge.org/mathematics
337 Methods in Banach space theory, J.M.F. CASTILLO & W.B. JOHNSON (eds)
338 Surveys in geometry and number theory, N. YOUNG (ed)
339 Groups St Andrews 2005 I, C.M. CAMPBELL, M.R. QUICK, E.F. ROBERTSON & G.C. SMITH (eds)
340 Groups St Andrews 2005 II, C.M. CAMPBELL, M.R. QUICK, E.F. ROBERTSON & G.C. SMITH (eds)
341 Ranks of elliptic curves and random matrix theory, J.B. CONREY, D.W. FARMER, F. MEZZADRI &
N.C. SNAITH (eds)
342 Elliptic cohomology, H.R. MILLER & D.C. RAVENEL (eds)
343 Algebraic cycles and motives I, J. NAGEL & C. PETERS (eds)
344 Algebraic cycles and motives II, J. NAGEL & C. PETERS (eds)
345 Algebraic and analytic geometry, A. NEEMAN
346 Surveys in combinatorics 2007, A. HILTON & J. TALBOT (eds)
347 Surveys in contemporary mathematics, N. YOUNG & Y. CHOI (eds)
348 Transcendental dynamics and complex analysis, P.J. RIPPON & G.M. STALLARD (eds)
349 Model theory with applications to algebra and analysis I, Z. CHATZIDAKIS, D. MACPHERSON,
A. PILLAY & A. WILKIE (eds)
350 Model theory with applications to algebra and analysis II, Z. CHATZIDAKIS, D. MACPHERSON,
A. PILLAY & A. WILKIE (eds)
351 Finite von Neumann algebras and masas, A.M. SINCLAIR & R.R. SMITH
352 Number theory and polynomials, J. MCKEE & C. SMYTH (eds)
353 Trends in stochastic analysis, J. BLATH, P. MÖRTERS & M. SCHEUTZOW (eds)
354 Groups and analysis, K. TENT (ed)
355 Non-equilibrium statistical mechanics and turbulence, J. CARDY, G. FALKOVICH & K. GAWEDZKI
356 Elliptic curves and big Galois representations, D. DELBOURGO
357 Algebraic theory of differential equations, M.A.H. MACCALLUM & A.V. MIKHAILOV (eds)
358 Geometric and cohomological methods in group theory, M.R. BRIDSON, P.H. KROPHOLLER &
I.J. LEARY (eds)
359 Moduli spaces and vector bundles, L. BRAMBILA-PAZ, S.B. BRADLOW, O. GARCÍA-PRADA &
S. RAMANAN (eds)
360 Zariski geometries, B. ZILBER
361 Words: Notes on verbal width in groups, D. SEGAL
362 Differential tensor algebras and their module categories, R. BAUTISTA, L. SALMERÓN & R. ZUAZUA
363 Foundations of computational mathematics, Hong Kong 2008, F. CUCKER, A. PINKUS & M.J. TODD (eds)
364 Partial differential equations and fluid mechanics, J.C. ROBINSON & J.L. RODRIGO (eds)
365 Surveys in combinatorics 2009, S. HUCZYNSKA, J.D. MITCHELL & C.M. RONEY-DOUGAL (eds)
366 Highly oscillatory problems, B. ENGQUIST, A. FOKAS, E. HAIRER & A. ISERLES (eds)
367 Random matrices: High dimensional phenomena, G. BLOWER
368 Geometry of Riemann surfaces, F.P. GARDINER, G. GONZÁLEZ-DIEZ & C. KOUROUNIOTIS (eds)
369 Epidemics and rumours in complex networks, M. DRAIEF & L. MASSOULIÉ
370 Theory of p-adic distributions, S. ALBEVERIO, A.YU. KHRENNIKOV & V.M. SHELKOVICH
371 Conformal fractals, F. PRZYTYCKI & M. URBAŃSKI
372 Moonshine: The first quarter century and beyond, J. LEPOWSKY, J. MCKAY & M.P. TUITE (eds)
373 Smoothness, regularity, and complete intersection, J. MAJADAS & A. G. RODICIO
374 Geometric analysis of hyperbolic differential equations: An introduction, S. ALINHAC
375 Triangulated categories, T. HOLM, P. JØRGENSEN & R. ROUQUIER (eds)
376 Permutation patterns, S. LINTON, N. RUŠKUC & V. VATTER (eds)
377 An introduction to Galois cohomology and its applications, G. BERHUY
378 Probability and mathematical genetics, N. H. BINGHAM & C. M. GOLDIE (eds)
379 Finite and algorithmic model theory, J. ESPARZA, C. MICHAUX & C. STEINHORN (eds)
380 Real and complex singularities, M. MANOEL, M.C. ROMERO FUSTER & C.T.C WALL (eds)
381 Symmetries and integrability of difference equations, D. LEVI, P. OLVER, Z. THOMOVA &
P. WINTERNITZ (eds)
382 Forcing with random variables and proof complexity, J. KRAJÍČEK
383 Motivic integration and its interactions with model theory and non-Archimedean geometry I, R. CLUCKERS,
J. NICAISE & J. SEBAG (eds)
384 Motivic integration and its interactions with model theory and non-Archimedean geometry II, R. CLUCKERS,
J. NICAISE & J. SEBAG (eds)
385 Entropy of hidden Markov processes and connections to dynamical systems, B. MARCUS, K. PETERSEN &
T. WEISSMAN (eds)
386 Independence-friendly logic, A.L. MANN, G. SANDU & M. SEVENSTER
387 Groups St Andrews 2009 in Bath I, C.M. CAMPBELL et al (eds)
388 Groups St Andrews 2009 in Bath II, C.M. CAMPBELL et al (eds)
389 Random fields on the sphere, D. MARINUCCI & G. PECCATI
390 Localization in periodic potentials, D.E. PELINOVSKY
391 Fusion systems in algebra and topology M. ASCHBACHER, R. KESSAR & B. OLIVER
392 Surveys in combinatorics 2011, R. CHAPMAN (ed)
393 Non-abelian fundamental groups and Iwasawa theory, J. COATES et al (eds)
394 Variational problems in differential geometry, R. BIELAWSKI, K. HOUSTON & M. SPEIGHT (eds)
Conference photograph
London Mathematical Society Lecture Note Series: 394
Variational Problems in
Differential Geometry
University of Leeds 2009
Edited by
R. BIELAWSKI
K. HOUSTON
J.M. SPEIGHT
University of Leeds
cambridge university press
Cambridge, New York, Melbourne, Madrid, Cape Town,
Singapore, São Paulo, Delhi, Tokyo, Mexico City
Cambridge University Press
The Edinburgh Building, Cambridge CB2 8RU, UK
Published in the United States of America by Cambridge University Press, New York
www.cambridge.org
Information on this title: www.cambridge.org/9780521282741
C Cambridge University Press 2012
A catalogue record for this publication is available from the British Library
v
vi Contents
Bernd Ammann
Facultät für Mathematik, Universität Regensburg, 93040 Regensburg,
Germany
Pierre Jammes
Laboratoire J.-A. Dieudonné, Université Nice – Sophia Antipolis, Parc
Valrose, F-06108 NICE Cedex 02, France
Claudio Arezzo
Abdus Salam International Center for Theoretical Physics, Strada Costiera
11, Trieste (Italy) and Dipartimento di Matematica, Università di Parma,
Parco Area delle Scienze 53/A, Parma, Italy
Gabriele La Nave
Department of Mathematics, Yeshiva University, 500 West 185 Street,
New York, NY, USA
Paul Baird
Département de Mathématiques, Université de Bretagne Occidentale,
6 Avenue Le Gorgeu – CS 93837, 29238 Brest, France
Josef F. Dorfmeister
Fakultät für Mathematik, Technische Universität München, Boltzmannstr. 3,
D-85747 Garching, Germany
viii
List of contributors ix
Akito Futaki
Department of Mathematics, Tokyo Institute of Technology, 2-12-1,
O-okayama, Meguro, Tokyo 152-8551, Japan
Yuji Sano
Department of Mathematics, Kyushu University, 6-10-1, Hakozaki,
Higashiku, Fukuoka-city, Fukuoka 812-8581 Japan
Frédéric Hélein
Institut de Mathématiques de Jussieu, UMR CNRS 7586, Université Denis
Diderot Paris 7, 175 rue du Chevaleret, 75013 Paris, France
Lorenz J. Schwachhöfer
Fakultät für Mathematik, Technische Universität Dortmund, Vogelpothsweg
87, 44221 Dortmund, Germany
Richard A. Wentworth
Department of Mathematics, University of Maryland, College Park, MD
20742, USA
Graeme Wilkin
Department of Mathematics, University of Colorado, Boulder, CO 80309,
USA
Jon Wolfson
Department of Mathematics, Michigan State University, East Lansing,
MI 48824, USA
Preface
xi
xii Preface
R. Bielawski
K. Houston
J.M. Speight
Leeds, UK
1
The supremum of first eigenvalues of
conformally covariant operators
in a conformal class
bernd ammann and pierre jammes
Abstract
Let (M, g) be a compact Riemannian manifold of dimension ≥ 3. We show that
there is a metric g̃ conformal to g and of volume 1 such that the first positive
eigenvalue of the conformal Laplacian with respect to g̃ is arbitrarily large.
A similar statement is proven for the first positive eigenvalue of the Dirac
operator on a spin manifold of dimension ≥ 2.
1.1 Introduction
The goal of this article is to prove the following theorems.
sup λ+
1 (Dg )Vol(M, g)
1/n
= ∞.
g∈[g0 ]
sup λ+
1 (Lg )Vol(M, g)
2/n
= ∞.
g∈[g0 ]
The Dirac operator and the conformal Laplacian belong to a large fam-
ily of operators, defined in details in subsection 1.2.3. These operators are
1
2 B. Ammann and P. Jammes
where λ1 (Lg ) is the lowest eigenvalue of Lg . If Y (M, [g0 ]) > 0, then the
solution of the Yamabe problem [29] tells us that the infimum is attained and
the minimizer is a metric of constant scalar curvature. This famous problem
was finally solved by Schoen and Yau using the positive mass theorem.
In a similar way, for n = 2 the Dirac operator is associated to constant-mean-
curvature conformal immersions of the universal covering into R3 . If a Dirac-
operator-analogue of the positive mass theorem holds for a given manifold
(M, g0 ), then the infimum
inf λ+
1 (Dg )Vol(M, g)
1/n
g∈[g0 ]
sup λ+
1 (Pg )Vol(M, g)
k/n
= ∞.
g∈[g0 ]
inf λ+
1 (Dg )Vol(M, g)
1/n
g∈[g0 ]
reflects a rich geometrical structure [3], [4], [5], [7], [8], similarly for the
conformal Laplacian. It seems natural to study the supremum as well.
The second motivation comes from comparing Theorem 1.1.3 to results
g
about some other differential operators. For the Hodge Laplacian p acting
g
on p-forms, we have supg∈[g0 ] λ1 (p )Vol(M, g) = +∞ for n ≥ 4 and 2 ≤
2/n
(the case k = 1 is proven in [20] and the general case in [27]). See [25] for a
synthetic presentation of this subject.
The essential idea in our proof is to construct metrics with longer and longer
cylindrical parts. We will call this an asymptotically cylindrical blowup. Such
metrics are also called Pinocchio metrics in [2, 6]. In [2, 6] the behavior of Dirac
eigenvalues on such metrics has already been studied partially, but the present
article has much stronger results. To extend these existing results provides the
third motivation.
Acknowledgments We thank B. Colbois, M. Dahl, and E. Humbert for
many related discussions. We thank R. Gover for some helpful comments on
conformally covariant operators, and for several references. The first author
4 B. Ammann and P. Jammes
wants to thank the Albert Einstein institute at Potsdam-Golm for its very kind
hospitality which enabled to write the article.
1.2 Preliminaries
1.2.1 Notations
In this article By (r) denotes the ball of radius r around y, Sy (r) = ∂By (r)
its boundary. The standard sphere S0 (1) ⊂ Rn in Rn is denoted by Sn−1 , its
volume is ωn−1 . For the volume element of (M, g) we use the notation dv g . In
our article, (V ) (resp. c (V )) always denotes the set of all smooth sections
(resp. all compactly supported smooth sections) of the vector bundle V → M.
For sections u of V → M over a Riemannian manifold (M, g) the Sobolev
norms L2 and H s , s ∈ N, are defined as
u 2L2 (M,g) := |u|2 dv g
M
u 2
H s (M,g) := u 2
L2 (M,g) + ∇u 2
L2 (M,g) + · · · + ∇s u 2
L2 (M,g) .
Pu = f (1.1)
on \ {0} with
−k
lim |u|r = 0 and lim |u| = 0 (1.2)
ε→0 B (2ε)−B (ε) ε→0 B (ε)
0 0 0
The supremum of first eigenvalues 5
Proof We show that u is a weak solution of (1.1) in the distributional sense, and
then it follows from standard regularity theory, that it is also a strong solution.
This means that we have to show that for any given compactly supported smooth
test function ψ : → R we have
uP ∗ ψ = f ψ.
We conclude
uP ∗ ψ = uP ∗ (ηψ) + uP ∗ ((1 − η)ψ)
= ∗
uP (ηψ) + (P u)(1 − η)ψ (1.4)
→0 → fψ
as in this case
2
−k 2 −k
|u|r ≤ |u| r r −k .
B0 (2ε)\B0 (ε) B0 (2ε)\B0 (ε)
≤C
6 B. Ammann and P. Jammes
Pg2 = f −
n+k n−k
2 Pg1 f 2 . (1.6)
One also expresses this by saying that P has bidegree ((n − k)/2, (n + k)/2).
The sense of this equation is apparent if Pg is an operator from C ∞ (M)
to C ∞ (M). If Pg acts on a vector bundle or if some additional structure (as
e.g. spin structure) is used for defining it, then a rigorous and careful defini-
tion needs more attention. The language of categories provides a good formal
framework [30]. The concept of conformally covariant elliptic operators is
already used by many authors, but we do not know of a reference where a
formal definition is carried out that fits to our context. (See [26] for a similar
categorial approach that includes some of the operators presented here.) Often
an intuitive definition is used. The intuitive definition is obviously sufficient if
one deals with operators acting on functions, such as the conformal Laplacian
or the Paneitz operator. However to properly state Theorem 1.1.3 we need the
following definition.
Let Riemn (resp. Riemspinn ) be the category n-dimensional Riemannian
manifolds (resp. n-dimensional Riemannian manifolds with orientation and
spin structure). Morphisms from (M1 , g1 ) to (M2 , g2 ) are conformal embed-
dings (M1 , g1 ) → (M2 , g2 ) (resp. conformal embeddings preserving orienta-
tion and spin structure).
Let Laplacenk (resp. Diracnk ) be the category whose objects are
{(M, g), Vg , Pg }
A morphism (ι, κ) from {(M1 , g1 ), Vg1 , Pg1 } to {(M2 , g2 ), Vg2 , Pg2 } consists
of a conformal embedding ι : (M1 , g1 ) → (M2 , g2 ) (preserving orientation
and spin structure in the case of Diracnk ) together with a fiber isomorphism
κ : ι∗ Vg2 → Vg1 preserving fiberwise length, such that Pg1 and Pg2 sat-
isfy the conformal covariance property (1.6). For stating this property pre-
cisely, let f > 0 be defined by ι∗ g2 = f 2 g1 , and let κ∗ : (Vg2 ) → (Vg1 ),
κ∗ (ϕ) = κ ◦ ϕ ◦ ι. Then the conformal covariance property is
κ∗ Pg2 = f −
n+k n−k
2 Pg1 f 2 κ∗ . (1.7)
In the following the maps κ and ι will often be evident from the context
and then will be omitted. The transformation formula (1.7) then simplifies
to (1.6).
g
For a fixed background metric g, the relation dC k (K) ( · , · ) defines a distance
function on the space of metrics on K. The topology induced by d g is inde-
pendent of this background metric and it is called the C k -topology of metrics
on K.
1.2.5 Examples
Example 1: The Conformal Laplacian
Let
n−2
Lg := g + Scalg ,
4(n − 1)
be the conformal Laplacian. It acts on functions on a Riemannian manifold
(M, g), i.e. Vg is the trivial real line bundle R. Let ι : (M1 , g1 ) → (M2 , g2 )
be a conformal embedding. Then we can choose κ := Id : ι∗ Vg2 → Vg1 and
formula (1.7) holds for k = 2 (see e.g. [15, Section 1.J]). All coefficients of
Lg depend continuously on g in the C 2 -topology. Hence L is a conformally
covariant elliptic operator of order 2 and of bidegree ((n − 2)/2, (n + 2)/2).
The scalar curvature of Sn−1 × R is (n − 1)(n − 2). The spectrum of Lg on
Sn−1 × R of Lg coincides with the essential spectrum of Lg and is [σL , ∞) with
σL := (n − 2)2 /4. Hence L is invertible on S n−1 × R if (and only if) n > 2.
Example 2: The Paneitz operator
Let (M, g) be a smooth, compact Riemannian manifold of dimension n ≥ 5.
The Paneitz operator Pg is given by
n−4
Pg u = (g )2 u − divg (Ag du) + Qg u
2
where
(n − 2)2 + 4 4
Ag := Scalg g − Ricg ,
2(n − 1)(n − 2) n−2
1 n3 − 4n2 + 16n − 16 2
Qg = g Scalg + Scal2g − |Ricg |2 .
2(n − 1) 8(n − 1) (n − 2)
2 2 (n − 2)2
The supremum of first eigenvalues 9
This operator was defined by Paneitz [32] in the case n = 4, and it was general-
ized by Branson in [17] to arbitrary dimensions ≥ 4. We also refer to Theorem
1.21 of the overview article [16]. The explicit formula presented above can
be found e.g. in [23]. The coefficients of Pg depend continuously on g in the
C 4 -topology
As in the previous example we can choose for κ the identity, and then the
Paneitz operator Pg is a conformally covariant elliptic operator of order 4 and
of bidegree ((n − 4)/2, (n + 4)/2).
On Sn−1 × R one calculates
(n − 4)n
Ag = Id + 4πR > 0
2
where πR is the projection to vectors parallel to R.
(n − 4)n2
Qg = .
8
We conclude
(n − 4)n2
σP = Q =
8
and P is invertible on Sn−1 × R if (and only if) n > 4.
Examples 3: The Dirac operator.
Let g̃ = f 2 g. Let g M resp. g̃ M be the spinor bundle of (M, g) resp.
g
(M, g̃). Then there is a fiberwise isomorphism βg̃ : g M → g̃ M, preserving
the norm such that
Dg̃ ϕ = f −
n+1 n−1
2 ◦ Dg f 2 ϕ . (1.8)
Locally
n−1
Dvert = ei · ∇ei
i=1
for a local frame (e1 , . . . , en−1 ) of Sn−1 . Here · denotes the Clifford multi-
plication T M ⊗ g M → g M. Furthermore Dhor = ∂t · ∇∂t , where t ∈ R is
the standard coordinate of R. The operators Dvert and Dhor anticommute. For
n ≥ 3, the spectrum of Dvert coincides with the spectrum of the Dirac operator
on Sn−1 , we cite [12] and obtain
n−1
specDvert = ± + k | k ∈ N0 .
2
The operator (Dhor )2 is the ordinary Laplacian on R and hence has spectrum
[0, ∞). Together this implies that the spectrum of the Dirac operator on Sn−1 ×
R is the set (−∞, −σD ] ∪ [σD , ∞) with σD = n−1 2
.
In the case n = 2 these statements are only correct if the circle Sn−1 = S1
carries the spin structure induced from the ball. Only this spin structure extends
to the conformal compactification that is given by adding one point at infinity
for each end. For this reason, we will understand in the whole article that all
circles S1 should be equipped with this bounding spin structure. The exten-
sion of the spin structure is essential in order to have a spinor bundle on the
compactification. The methods used in our proof use this extension implicitly.
Hence D is invertible on S n−1 × R if (and only if) n > 1.
Most techniques used in the literature on estimating eigenvalues of the
Dirac operators do not use the spin structure and hence these techniques cannot
provide a proof in the case n = 2.
Example 4: The Rarita-Schwinger operator and many other Fegan type
operators are conformally covariant elliptic operators of order 1 and of bide-
gree ((n − 1)/2, (n + 1)/2). See [21] and in the work of T. Branson for more
information.
Example 5: Assume that (M, g) is a Riemannian spin manifold that carries
a vector bundle W → M with metric and metric connection. Then there is a
natural first order operator (g M ⊗ W ) → (g M ⊗ W ), the Dirac opera-
tor twisted by W . This operator has similar properties as conformally covariant
elliptic operators of order 1 and of bidegree ((n − 1)/2, (n + 1)/2). The meth-
ods of our article can be easily adapted in order to show that Theorem 1.1.3
is also true for this twisted Dirac operator. However, twisted Dirac operators
are not “conformally covariant elliptic operators” in the above sense. They
could have been included in this class by replacing the category Riemspinn by
The supremum of first eigenvalues 11
. . . ≤ λ− + +
1 (PL ) < 0 = 0 . . . = 0 < λ1 (PL ) ≤ λ2 (PL ) ≤ . . . ,
for all u ∈ H k (M∞ ,g∞ ) (V ) and all s ∈ {0, 1, . . . , k}. The operator
ϕ → f −
n−k
2 ϕ
obviously defines an isomorphism from ker Pg to ker Pg̃ . It is less obvious that
a similar statement holds if we compare g0 and g∞ defined before:
ker P0 → ker P∞
− n−k
ϕ0 → ϕ∞ = F∞ 2
ϕ0
Here we used that up to lower order terms dv g∞ coincides with the product
measure of the standard measure on the sphere with the measure d(log r) =
1
r
dr. Furthermore, formula (1.6) implies P∞ ϕ∞ = 0. Hence the map is well-
defined. In order to show that it is an isomorphism we show that the obvious
n−k
inverse ϕ∞ → ϕ0 := F∞2 ϕ∞ is well defined. To see this we start with an
L2 -section in the kernel of P∞ .
14 B. Ammann and P. Jammes
We calculate
k
F∞ |ϕ0 |2 dv g0
= |ϕ∞ |2 dv g∞ .
M M∞
Using again (1.6) we see that this section satisfies P0 ϕ0 = 0 on M \ {y}. Hence
condition (1.5) is satisfied, and together with the removal of singularity lemma
(Lemma 1.2.1) one obtains that the inverse map is well defined. The proposition
follows.
λ± ±
j (PL ) → λj (P∞ ) ∈ (−σP , σP ) for L → ∞.
In the case Spec(Pg0 ) ⊂ (0, ∞) the theorem only makes a statement about
λ+j and conversely in the case that Spec(Pg0 ) ⊂ (−∞, 0) it only makes a
,
statement about λ−
j .
Obviously this theorem implies Theorem 1.1.3.
lim sup(λ+ +
j (PL )) ≤ λj (P∞ ). (1.11)
L→∞
lim sup(−λ− −
j (PL )) ≤ −λj (P∞ ). (1.12)
L→∞
(for n > k) the norm ψi, L2 (M∞ ,g∞ ) is finite as well, and we can renormalize
such that
Lemma 1.4.2 For any δ > 0 and any ∈ {0, . . . , m} the sequence
ψi, C k+1 (M\By (δ),g∞ ) i
is bounded.
Proof of the lemma. After removing finitely many i, we can assume that λi ≤
2λ̄ and e−Li < δ/2. Hence FL = F∞ and hi = λi on M \ By (δ/2). Because of
|(P∞ ) ψi | dv
s 2 g∞
≤ (2λ̄) 2s
|ψi |2 dv g∞ ≤ (2λ̄)2s
M\By (δ/2) M\By (δ/2)
k,α
By taking a diagonal sequence, one can obtain convergence in Cloc (M∞ ) of
ψi, to ψ̄ . It remains to prove that ψ̄1 , . . . ,ψ̄m are linearly independent, in
particular that any ψ̄ = 0. For this we use the following lemma.
we get
for all s ∈ {0, . . . , k}. Let χ be a cut-off function as in Subsection 1.4.2 with
T = − log δ. Hence
C C
P∞ (1 − χ )ψi, − (1 − χ )P∞ (ψi, ) L2 (M∞ ,g∞ ) ≤ = . (1.14)
T − log δ
On the other hand (By (δ) \ {y}, g∞ ) converges for suitable choices of base
points for δ → 0 to Sn−1 × (0, ∞) in the C ∞ -topology of Riemannian man-
ifolds with base points. Hence there is a function τ (δ) converging to 0 such
that
P∞ (1 − χ )ψi, L2 (M∞ ,g∞ ) ≥ (σp − τ (δ)) (1 − χ )ψi, L2 (M∞ ,g∞ ) . (1.15)
The right hand side is smaller than ε for i sufficiently large and δ suffi-
ciently small. The main statement of the lemma then follows for δ0 := δ 2 .
The Minkowski inequality yields.
and thus ψ̄ L2 (M∞ ,g∞ ) = 1. The orthogonality of these sections is pro-
vided by the following lemma, and the inequality (1.13) then follows
immediately.
Proof of the lemma. The sections ϕi,1 , . . . , ϕi, are orthogonal. For any fixed
δ0 (given by the previous lemma), it follows for sufficiently large i that
ψi, , ψi,˜ dv g∞ = ϕi, , ϕi,˜ dv gLi
M\By (δ0 ) M\By (δ0 )
= ϕi, , ϕi,˜ dv gLi
By (δ0 )
FLi k (1.16)
= ψi, , ψi,˜ dv g∞
By (δ0 ) F∞
≤1
≤ ε2
Because of strong L2 convergence on M \ By (δ0 ) this implies
ψ̄ , ψ̄˜ dv g∞ ≤ ε2 (1.17)
M\By (δ0 )
The aim of this appendix is to sketch how to prove Proposition 1.3.1. All
properties in this appendix are well-known to experts, but explicit references
are not evident to find. Thus this summary might be helpful to the reader.
The geometry of (M∞ , g∞ ) is asymptotically cylindrical. The metric g∞
is even a b-metric in the sense of Melrose [31], but to keep the presentation
simple, we avoid the b-calculus.
If (r, γ ) ∈ R+ × Sn−1 denote polar normal coordinates with respect to the
metric g0 , and if we set t := − log r, then (t, γ ) defines a diffeomorphism α :
By 0 (1/2) \ {y} → [log 2, ∞) × Sn−1 such that (α −1 )∗ g∞ = dt 2 + ht for a
(M,g )
family of metrics such that (α −1 )∗ g∞ , all of its derivatives, its curvature, and all
derivatives of the curvature tend to the standard metric on the cylinder, and the
speed of the convergence is majorised by a multiple of et . Thus the continuity
of the coefficients property implies, that P∞ extends to a bounded operator
from H k (M∞ ,g∞ ) (V ) → L2 (M∞ ,g∞ ) (V ).
The formal self-adjointness of P∞ implies that
ψ, P∞ ϕ = P∞ ψ, ϕ (A.18)
M∞ M∞
19
20 B. Ammann and P. Jammes
See also [28, III §3] for a good presentation on how to construct and work with
such a parametrix.
To see the uniformicity, one verifies that
u H k (K,g̃)
− 1 ≤ C g̃ − g C k ≤ Cε
u k
H (K,g)
and
P̃ (u)
L2 (U )
− 1 ≤ Cε u H k (U ) .
P (u)
L2 (U )
ϕ ∈ H k (M∞ ,g∞ ) (V ).
References
[1] B. Ammann, The Dirac Operator on Collapsing Circle Bundles, Sém. Th. Spec.
Géom Inst. Fourier Grenoble 16 (1998), 33–42.
[2] B. Ammann, Spin-Strukturen und das Spektrum des Dirac-Operators, Ph.D. thesis,
University of Freiburg, Germany, 1998, Shaker-Verlag Aachen 1998, ISBN 3-
8265-4282-7.
[3] , The smallest Dirac eigenvalue in a spin-conformal class and cmc-
immersions, Comm. Anal. Geom. 17 (2009), 429–479.
[4] , A spin-conformal lower bound of the first positive Dirac eigenvalue, Diff.
Geom. Appl. 18 (2003), 21–32.
[5] , A variational problem in conformal spin geometry, Habilitationsschrift,
Universität Hamburg, 2003.
[6] B. Ammann and C. Bär, Dirac eigenvalues and total scalar curvature, J. Geom.
Phys. 33 (2000), 229–234.
[7] B. Ammann and E. Humbert, The first conformal Dirac eigenvalue on 2-
dimensional tori, J. Geom. Phys. 56 (2006), 623–642.
[8] B. Ammann, E. Humbert, and B. Morel, Mass endomorphism and spinorial Yam-
abe type problems, Comm. Anal. Geom. 14 (2006), 163–182.
[9] B. Ammann, A. D. Ionescu, and V. Nistor, Sobolev spaces on Lie manifolds and
regularity for polyhedral domains, Doc. Math. 11 (2006), 161–206.
[10] B. Ammann, R. Lauter, and V. Nistor, On the geometry of Riemannian manifolds
with a Lie structure at infinity, Int. J. Math. Math. Sci. (2004), 161–193.
[11] , Pseudodifferential operators on manifolds with a Lie structure at infinity,
Ann. of Math. 165 (2007), 717–747.
[12] C. Bär, The Dirac operator on space forms of positive curvature, J. Math. Soc.
Japan 48 (1996), 69–83.
[13] C. Bär, The Dirac operator on hyperbolic manifolds of finite volume, J. Differ.
Geom. 54 (2000), 439–488.
[14] H. Baum, Spin-Strukturen und Dirac-Operatoren über pseudoriemannschen Man-
nigfaltigkeiten, Teubner Verlag, 1981.
[15] A. L. Besse, Einstein manifolds, Ergebnisse der Mathematik und ihrer Grenzgebi-
ete, 3. Folge, no. 10, Springer-Verlag, 1987.
[16] T. P. Branson, Differential operators canonically associated to a conformal struc-
ture, Math. Scand. 57 (1985), 293–345.
[17] , Group representations arising from Lorentz conformal geometry, J. Funct.
Anal. 74 (1987), no. 2, 199–291.
[18] , Second order conformal covariants, Proc. Amer. Math. Soc. 126 (1998),
1031–1042.
[19] B. Colbois and A. El Soufi, Eigenvalues of the Laplacian acting on p-forms and
metric conformal deformations, Proc. of Am. Math. Soc. 134 (2006), 715–721.
[20] A. El Soufi and S. Ilias, Immersions minimales, première valeur propre du laplacien
et volume conforme, Math. Ann. 275 (1986), 257–267.
[21] H. D. Fegan, Conformally invariant first order differential operators., Quart. J.
Math. Oxford, II. series 27 (1976), 371–378.
[22] R. Gover and L. J. Peterson, Conformally invariant powers of the Laplacian,
Q-curvature, and tractor calculus, Comm. Math. Phys. 235 (2003), 339–378.
The supremum of first eigenvalues 23
[23] E. Hebey and F. Robert, Coercivity and Struwe’s compactness for Paneitz type
operators with constant coefficients, Calc. Var. Partial Differential Equations 13
(2001), 491–517.
[24] N. Hitchin, Harmonic spinors, Adv. Math. 14 (1974), 1–55.
[25] P. Jammes, Extrema de valeurs propres dans une classe conforme, Sémin. Théor.
Spectr. Géom. 24 (2007), 23–42.
[26] I. Kolář, P. W. Michor, and J. Slovák, Natural operations in differential geometry,
Springer-Verlag, Berlin, 1993.
[27] N. Korevaar, Upper bounds for eigenvalues of conformal metrics, J. Differ. Geom.
37 (1993), 73–93.
[28] H.-B. Lawson and M.-L. Michelsohn, Spin Geometry, Princeton University Press,
1989.
[29] J. M. Lee and T. H. Parker. The Yamabe problem. Bull. Am. Math. Soc., New Ser.
17 (1987), 37–91.
[30] S. Mac Lane, Categories for the working mathematician, Graduate Texts in Math-
ematics, vol. 5, Springer-Verlag, New York, 1998.
[31] R. B. Melrose, The Atiyah-Patodi-Singer index theorem, Research Notes in Math-
ematics, vol. 4, A K Peters Ltd., Wellesley, MA, 1993.
[32] S. M. Paneitz, A quartic conformally covariant differential operator for arbitrary
pseudo-Riemannian manifolds, Preprint 1983, published in SIGMA 4 (2008).
[33] M. E. Taylor, M. E., Pseudodifferential operators, Princeton University Press,
1981.
Authors’ addresses:
Bernd Ammann
Facultät für Mathematik
Universität Regensburg
93040 Regensburg
Germany
[email protected]
Pierre Jammes
Laboratoire J.-A. Diendonné, Université Nice-Sophia Antipolis,
Parc Valrose, F-06108 Nice Cedex02, France
[email protected]
2
K-Destabilizing test configurations with
smooth central fiber
claudio arezzo, alberto della vedova, and
gabriele la nave
Abstract
In this note we point out a simple application of a result by the authors in
[2]. We show how to construct many families of strictly K-unstable polarized
manifolds, destabilized by test configurations with smooth central fiber. The
effect of resolving singularities of the central fiber of a given test configuration
is studied, providing many new examples of manifolds which do not admit
Kähler constant scalar curvature metrics in some classes.
2.1 Introduction
In this note we want to speculate about the following Conjecture due to Tian-
Yau-Donaldson ([23], [24], [25], [7]):
The notion of K-stability will be recalled below. For the moment it suffices to
say, loosely speaking, that a polarized manifold, or more generally a polarized
variety (V , A), is K-stable if and only if any special degeneration or test
configuration of (V , A) has an associated non positive weight, called Futaki
invariant and that this is zero only for the product configuration, i.e. the trivial
degeneration.
We do not even attempt to give a survey of results about Conjecture 2.1.1, but
as far as the results of this note are concerned, it is important to recall the reader
that Tian [24], Donaldson [7], Stoppa [22], using the results in [3] and [4], and
Mabuchi [17] have proved the sufficiency part of the Conjecture. Destabilizing
a polarizing manifold then implies non existence results of Kähler constant
scalar curvature metrics in the corresponding classes.
24
K-Destabilizing test configurations 25
One of the main problems in this subject is that under a special degeneration
a smooth manifold often becomes very singular, in fact just a polarized scheme
in general. This makes all the analytic tool available at present very difficult to
use.
Hence one naturally asks which type of singularities must be introduced to
make the least effort to destabilize a smooth manifold without cscK metrics.
The aim of this note is to provide a large class of examples of special
degenerations with positive Futaki invariant and smooth limit. In fact we want to
provide a “machine” which associates to any special degeneration of a polarized
normal variety (V , A) with positive Futaki invariant a special degeneration for
a polarized manifold (M̃, Ã) with smooth central fiber and still positive Futaki
invariant.
To the best of our knowledge, before this work the only known examples of
special degeneration with non negative Futaki invariant and smooth central fiber
are the celebrated example of Mukai-Umemura’s Fano threefold ([18]) used
by Tian in [24] to exhibit the first examples of Fano manifolds with discrete
automorphism group and no Kähler-Einstein metrics (other Fano manifolds
with these properties have been then produced in [1]). In this case there exist
non trivial special degenerations with smooth limit and zero Futaki invariant
(hence violating the definition of K-stability). It then falls in the borderline
case, making this example extremely interesting and delicate. We stress that
our “machine” does not work in this borderline case, because a priori the Futaki
invariant of the new test configuration is certainly small (by [2]) but we cannot
control its sign.
To state our result more precisely we now recall the relevant definitions:
usual weight one C∗ -action) such that L|f −1 (0) is ample on f −1 (0) and we have
(f −1 (1), L|f −1 (1) ) (V , Ar ) for some r > 0.
When (V , A) has a C∗ -action ρ : C∗ → Aut(V ), a test configuration where
X = V × C and C∗ acts on X diagonally through ρ is called product configu-
ration.
Given a test configuration (X, L) we will denote by F (X, L) the Futaki
invariant of the C∗ -action induced on the central fiber (f −1 (0), L|f −1 (0) ).
If (X, L) is a product configuration as above, clearly we have F (X, L) =
F (V , A, ρ).
Definition 2.1.4 The polarized manifold (M, A) is K-stable if for each test
configuration for (M, A) the Futaki invariant of the induced action on the central
fiber (f −1 (0), L|f −1 (0) ) is less than or equal to zero, with equality if and only if
we have a product configuration.
A test configuration (X, L) is called destabilizing if the Futaki invariant of
the induced action on (f −1 (0), L|f −1 (0) ) is greater than zero.
Test configurations for an embedded variety V ⊂ PN endowed with the hyper-
plane polarization A can be constructed as follows. Given a one-parameter
subgroup ρ : C∗ → GL(N + 1), which induces an obvious diagonal C∗ -action
on PN × C, it clear that the subscheme
X = (z, t) ∈ PN × C | t = 0, (ρ(t −1 )z, t) ∈ V ⊂ PN × C,
is invariant and projects equivariantly on C. Thus considering the relatively
ample polarization L induced by the hyperplane bundle gives test configuration
for (V , A). On the other hand, given a test configuration (X, L) for a polarized
variety (V , A), the relative projective embedding given by Lr , with r sufficiently
large, realizes X as above (see details in [21]).
We can now describe our “machine”: consider a test configuration (X, L)
for a polarized normal variety (V , A) with F (X, L) > 0. Up to raise L to a
suitable power – which does not affect the Futaki invariant – we can suppose
being in the situation above with X ⊂ PN × C invariantly, and L induced by the
hyperplane bundle of PN . At this point we consider the central fiber X0 ⊂ PN ,
which is invariant with respect to ρ, and we apply the (equivariant) resolution of
singularities [14, Corollary 3.22 and Proposition 3.9.1]. Thus there is a smooth
manifold P̃ acted on by C∗ and an equivariant map
β : P̃ → PN
which factorizes through a sequence of blow-ups, such that the strict transform
X̃0 of X0 is invariant and smooth. The key observation is that the strict transform
X̃1 of the fiber X1 ⊂ X degenerate to X0 under the given C∗ action on P̃ , thus it
K-Destabilizing test configurations 27
π : X̃ → X.
The issue raised at point (2) was addressed in [2] and it was proved that the
following natural (topological) definition makes the Futaki invariant a continu-
ous function around big and nef points in the Kähler cone. We will give simple
self-contained proofs in the cases of smooth manifolds and varieties with just
normal singularities in Section 2.
w(V , B k )
= F0 k + F1 + O(k −1 ),
χ (V , B k )
and we define
F (V , B) = F1
The existence of the expansion involved in definition above follows from the
standard fact that χ (V , B k ) is a polynomial of degree dim V , whose proof (see
for example [11]) can be easily adapted to show that w(V , B k ) is a polynomial
of degree at most dim V + 1.
28 C. Arezzo, A. Della Vedova, and Gabriele La Nave
1
F (V , B r ⊗ A) = F (V , B) + O , as r → ∞.
r
Having established a good continuity property of the Futaki invariant up to these
boundary point, we need to address the question of the effect of a resolution
of singularities of the central fiber. This is a particular case of the following
non trivial extension of previous analysis by Ross and Thomas [21] which was
proved in [2] where the general case of birational morphisms has been studied:
with strict inequality if and only if the support of β∗ (OX )/OX has codimension
one.
The proof of these results uses some heavy algebraic machinery, yet their proof
when (V , A) or the central fiber of (X, L) have only normal singularities (a
case largely studied) is quite simple and we give it in Section 2.
The Corollary of Theorem 2.1.6 and Theorem 2.1.7 we want to point out in
this note is then the following:
Theorem 2.1.8 Let (X, L) be a test configuration for the polarized normal
variety (V , A) with positive Futaki invariant. Let moreover (X̃, L̃) be a (big
and nef) test configuration obtained from (X, L) as above and let (M̃, B̃) be
the smooth (big and nef) fiber over the point 1 ∈ C. Let R be any relatively
ample line bundle over X̃.
Then (X̃, L̃r ⊗ R) is a test configuration for (M̃, B̃ r ⊗ R|M̃ ) with following
properties:
While this Theorem clearly follows from Theorems 2.1.6 and Theorem 2.1.7,
but for the specific case of central fiber with normal singularities it follows
from the much simpler Proposition 2.2.1 and Theorem 2.2.3.
The range of applicability of the above theorem is very large. We go through
the steps of the resolution of singularities in an explicit example by Ding-
Tian [6] of a complex orbifold of dimension 2. In this simple example explicit
calculations are easy to perform, yet we point out that the final example is
somehow trivial since it ends on a product test configuration. On the other
hand abundance of similar examples even in dimension 2 can be obtained by
the reader as an exercise using the results of Jeffres [12] and Nagakawa [19],
in which cases we loose an explicit description of the resulting destabilized
manifold, but we get new nontrivial examples. In fact in higher dimensions one
can use the approach described in this note to test also the Arezzo-Pacard blow
up theorems [3] [4], when the resolution of singularities requires a blow up of
a scheme of positive dimension.
Proposition 2.2.1 Let A, L be respectively an ample and a big and nef line
bundle on a smooth projective manifold M. For every C∗ -action on M that
linearizes to A and L, as r → +∞ we have
1
F (M, Lr ⊗ A) = F (M, L) + O .
r
Proof The result is a simple application of the equivariant Riemann-Roch
Theorem. We present here the details of the calculations involved, since we
could not find precise references for them.
Fix an hermitian metrics on A that is invariant with respect to the action of
S 1 ⊂ C∗ and suppose that the curvature ω is a Kähler metric. Since L is nef,
for each r > 0 we can choose an invariant metric on L whose curvature ηr
satisfy rηr + ω > 0. In other words rηr + ω is a Kähler form which coincides
with the curvature of the induced hermitian metric on the line bundle Lr ⊗ A.
30 C. Arezzo, A. Della Vedova, and Gabriele La Nave
As r → +∞ we have expansions:
(ηr + 1r ω)n c1 (Lr ⊗ A)n c1 (L)n 1
= n
= +O
M n! n! r n! r
(ηr + 1
ω)n−1 ∧ Ric(ηr + 1r ω) c1 (Lr ⊗ A)n−1 c1 (M) 1
r
= +O
M 2(n − 1)! 2(n − 1)! r n−1 r
c1 (L)n−1 c1 (M) 1
= +O
2(n − 1)! r
1 (ηr + 1r ω)n cT (Lr ⊗ A)n+1
gr + f = 1
M r n! (n + 1)! r n+1
c1T (L)n+1 1
= +O
(n + 1)! r
1 (ηr + 1r ω)n−1 ∧ Ric(ηr + 1r ω) cT (Lr ⊗ A)n c1T (M)
gr + f = 1
M r 2(n − 1)! 2 n! r n
c1T (L)n c1T (M) 1
= +O ,
2 n! r
K-Destabilizing test configurations 31
where c1T denotes the equivariant first Chern class. Thus we have:
c1T (L)n c1T (M) c1 (L)n c1T (L)n+1 c1 (L)n−1 c1 (M)
2n!
· n!
− (n+1)!
· 2(n−1)! 1
F (M, L ⊗ A) =
r
c1 (L)n 2 +O .
r
n!
and the thesis follows applying again the (equivariant) Riemann-Roch theorem
to L.
The following has essentially been proved by Paul-Tian ([20]); here we give a
simple self-contained proof.
F (V , L) = F (Ṽ , π ∗ L).
Combining Propositions 2.2.1 and 2.2.2 we get the following result, which
explains the behaviour for normal varieties of the Futaki invariant under Hiron-
aka’s resolution of singularities process:
where the second and third equalities follow from Propositions 2.2.1 and 2.2.2
respectively.
1
F (X̃, L̃r ⊗ R) = F (X̃, L̃) + O . (2.3)
r
Let us collect some elementary facts that the reader can easily verify:
1 Xf is singular only at p0 = (1 : 0 : 0 : 0).
2 Having set (x, y, z) affine co-ordinates centered on (1 : 0 : 0 : 0), let
C = {((x, y, z), (l0 : l1 : l2 )) ∈ C3 × P2 |
xl1 − yl0 = xl2 − zl0 = yl2 − zl1 = 0}
be the blow-up of C3 at the origin with exceptional divisor E P2 and Xf
the proper transform of Xf , then Xf is singular at points
(0 : 1 : 0), (0 : 0 : 1), (0 : 1 : 1) ∈ E.
3 Having set X̃f the proper transform of Xf under the blow-up C̃ of C at
points (0 : 1 : 0), (0 : 0 : 1), (0 : 1 : 1) ∈ E and E1 the exceptional divisor
over (0 : 1 : 0), then X̃f is smooth around E1 . Analogously, if E2 is the
exceptional divisor over (0 : 0 : 1) and E3 the one over (0 : 1 : 1), then X̃f
is smooth around E2 and E3 too. For future reference let E0 be the proper
transform of E under the second blow up.
4 X̃f is smooth.
5 Consider now the C∗ -action on P3 defined by
t · (z0 , z1 , z2 , z3 ) = (t α0 z0 , t α1 z1 , t α2 z2 , t α3 z3 )
with αj ∈ Z and α0 + · · · + α3 = 0, thus
(t · f )(z) = t −(α0 +2α1 ) z0 z12 + t −(2α2 +α3 ) z22 z3 − t −(α2 +2α3 ) z2 z32 ,
and f is semi-invariant if and only if
α0 + 2α1 = 2α2 + α3 = α2 + 2α3 ,
hence
(α0 , . . . , α3 ) = (−7β, 5β, β, β),
with β ∈ Z, and f has weight −3β.
6 The monomials of degree three in the variables z0 , . . . , z3 are a basis of
semi-invariants for the fixed C∗ -action on C[z0 , . . . , z3 ]3 . In particular fix
β = −1, then the subspace spanned by monomials with weight greater or
equal to 4 is
V = span{z13 , z12 z2 , z12 z3 , z1 z22 , z1 z2 z3 , z1 z32 }
Thus the hypersurfaces that degenerate to Xf under the fixed C∗ -action (with
β = −1) are of the form
Xg = {f + g = 0},
where g ∈ V .
34 C. Arezzo, A. Della Vedova, and Gabriele La Nave
Thus X̃f is the central fiber of the special degeneration of X̃g given by the chosen
C∗ -action. Denoted byπ4 : P̃ → P the blow-up map, let Ar be the restriction
3
∗
of π OP3 (r) ⊗ O − =0 Ej to X̃f . By Theorem 2.2.3, F (X̃f , Ar ) has the
same sign of F (Xf , OXf (1)) when r is large enough. But thanks to [6] (see also
[16]), with our sign convention, (Xf , OXf (1)) has positive Futaki invariant, then
X̃g polarized with the restriction of π ∗ OP3 (r) ⊗ O − 4=0 Ej is K-unstable.
References
[1] C. Arezzo, A. Della Vedova, On the K-stability of complete intersections in polar-
ized manifolds, arXiv:0810.1473.
[2] C. Arezzo, A. Della Vedova and G. La Nave, Singularities and K-semistability,
arXiv:0906.2475.
[3] C. Arezzo and F. Pacard, Blowing up and desingularizing Kähler orbifolds with
constant scalar curvature, Acta Mathematica 196, no 2, (2006) 179-228.
[4] C. Arezzo and F. Pacard, Blowing up Kähler manifolds with constant scalar
curvature II, Annals of Math. 170 no. 2, (2009) 685-738
[5] D. Eisenbud, J. Harris, The Geometry of schemes. GTM 197. Springer, 2000.
[6] W. Y. Ding and G. Tian, Kähler-Einstein metrics and the generalized Futaki
invariant. Invent. Math. 110 (1992), no. 2, 315–335.
[7] S. K. Donaldson, Scalar curvature and stability of toric varieties. J. Differential
Geom. 62 (2002), no.2, 289–349.
[8] S. K. Donaldson. Lower bounds on the Calabi functional, J. Differential Geom.,
70 (2005), no.3, 453–472.
[9] J. Fine and J. Ross, A note on positivity of the CM line bundle. Int. Math. Res.
Not., 2006.
[10] A. Futaki and T. Mabuchi, Bilinear forms and extremal Kähler vector fields asso-
ciated with Kähler classes. Math. Ann. 301 (1995), n.2, 199–210
[11] R. Hartshorne, Algebraic Geometry, Springer, 1977.
[12] T. Jeffres, Singular set of some Kähler orbifolds. Trans. Amer. Math. Soc. 349
(1997), no. 5, 1961–1971.
[13] G. Kempf, F. Knudsen, D. Mumford and B. Saint-Donat, Toroidal Embeddings I
Lecture Notes in Mathematics, 339. Springer, 1973.
[14] J. Kollár, Lectures on resolution of singularities. Annals of Mathematics Studies,
166. Princeton University Press, Princeton, NJ, 2007.
[15] R. Lazarsfeld. Positivity in Algebraic Geometry I. Springer, 2004.
[16] Z. Lu, On the Futaki invariants of complete intersections. Duke Math. J. 100
(1999), no.2, 359–372.
K-Destabilizing test configurations 35
Authors’ addresses:
Claudio Arezzo
Abdus Salam International Center for Theoretical Physics,
Strada Costiera 11,
Trieste (Italy) and Dipartimento di Matematica,
Università di Parma,
Parco Area delle Scienze 53/A,
Parma, Italy
[email protected]
Gabriele La Nave
Department of Mathematics
Yeshiva University
500 West 185 Street
New York, NY
USA
[email protected]
3
Explicit constructions of Ricci solitons
paul baird
Abstract
We describe methods for constructing explicit examples of Ricci solitons. An
improved version of an anzatz of the author and L. Danielo for 3-dimensional
solitons is given which we extend to higher dimensions. The soliton structure
on the 4-dimensional geometry Nil4 is analysed in detail, in particular its
uniqueness is established and its relation to the natural Riemannian projection
Nil4 → Nil3 is discussed.
3.1 Introduction
The Ricci flow is the evolution equation:
∂g
= −2Ricci (g) (3.1)
∂t
for a time-dependent Riemannian metric g = g(t) defined on an n-dimensional
manifold M n , subject to some initial condition g(0) = g0 . This equation has
received a considerable amount of attention recently due to the work of G.
Perelman [14] which has advanced Hamilton’s program for resolving the
geometrization conjecture. An essential part of this story are Ricci solitons,
which correspond to fixed points, up to scaling and diffeomorphism. Specifi-
cally, if g(t) = c(t)ψt∗ (g0 ) is a solution of (3.1) on some time interval [0, δ),
where c(t) is a family of positive scalars such that c(0) = 1 and ψt is a family
of diffeomorphisms satisfying ψ0 = id, then on calculating g (0) and applying
(3.1) we obtain the equation:
37
38 P. Baird
where we now write g = g0 , and where 2A = c (0) and E is the vector field
(called the soliton flow) determined at each x ∈ M by E(x) = dtd ψt (x)|t=0 .
Conversely, on a complete manifold, any solution g0 of (3.2) determines
a solution of (3.1) of the form g(t) = c(t)ψt∗ (g0 ) for some small time t
(see [5]).
Equation (3.2) is known as the soliton equation and solutions play a fun-
damental role in the study of the Ricci flow: they occur as rescaled limits at
singularity formation and as asymptotic limits of immortal solutions, that is
solutions that exist for all future time [13]. Both of these limits are interpreted
in terms of Cheeger-Gromov-Hamilton pointed convergence of Ricci flows [9].
In the case when E = grad f is the gradient of a function, then the soliton is
said to be of gradient type, in which case (3.2) becomes
fixed points gives a complete metric of positive sectional curvature [4] (see Sec-
tion 3.4). This procedure was generalized by T. Ivey to construct solitons from
doubly warped product metrics of the type dr 2 + a(r)2 g1 + b(r)2 g2 , where
g1 and g2 are metrics on a sphere and an Einstein manifold, resp. [11]. In
Section 3.2, we make a further generalization which leads to a dynamical sys-
tem. Here we no longer insist that the soliton be of gradient type and neither
do we require that the soliton flow depend on a single parameter. The proce-
dure picks up solitons of Sol-type, as well as a curious example on a surface.
Although the soliton metric is neither complete nor compact, it factors and then
extends by the addition of two points to a non-standard (complete) metric on
the 2-sphere (Example 3.2.6).
In [2], the author and L. Danielo devised a procedure for constructing
solitons in dimension 3, by reducing the number of variables to two, so that
(3.2) becomes a system of equations on a Riemannian surface. It is important
to note that the method does not consist simply of supposing quantities depend
on two independent variables, rather, having solved the system of equations
on the surface one is then required to solve an exterior differential equation in
dimension 3 in order to obtain the 3-dimensional soliton metric. In particular,
the metric itself and the soliton flow may depend on all three coordinates. In
Section 3.3, we give an improved version of this ansatz and describe how
the soliton structure on Nil3 arises. Then in Section 3.4, we generalize the
construction to more general dimensions.
On Rn there are the well-known Gaussian solitons with flow E =
grad(− A2 |x − a|2 ), where a ∈ Rn is some arbitrary point and A is an arbi-
trary constant. This shows that a soliton structure may not be unique. However,
it is not known whether there are other examples of Riemannian manifolds
which support more than one soliton structure. One way to test this for a given
manifold, is to solve the equation (3.2) directly. This requires sufficient sym-
metry of the metric in order that the system of equations is managable. In [2],
this was done for the 3-dimensional geometries and uniqueness was established
in all cases except R3 , where one readily sees that the only examples are the
Gaussian solitons. In Section 3.5, we do this for the 4-dimensional geometry
Nil4 , to obtain the soliton structure whose existence was demonstrated by J.
Lauret [12] by Lie group methods and is described more explicitly by J. Lott
[13]. We further show uniqueness of this structure and discuss its relation to
the one on Nil3 : the geometry Nil4 naturally admits a harmonic Riemannian
fibration to Nil3 , which allows us to relate their respective soliton structures.
The author thanks the Agence Nationale de Recherche project: Flots et
opérateurs géométriques no: ANR-07-BLAN-0251-01 for financial support
during the preparation of this paper.
40 P. Baird
where the functions a, b both depend on x 1 only. As we shall see, the vector
field E defining the soliton flow may also depend on the variable x 3 . Then the
components of the Ricci curvature are given by the formula:
Rj k = ∂l jl k − ∂j lk
l
+ lm
l
jmk − jl m lk
m
,
and we find that the Ricci curvature is also diagonal with components:
E = X + fU (3.5)
Example 3.2.2 (The geometry Sol) For A > 0, the equilibrium points are
given by
√ √ √ !
±k A ∓A A k 2 + A2
√ ,√ ,± √ .
k 2 + A2 k 2 + A2 A
This gives the metric
⎛ ⎞
1 0
#
0
⎜ A ⎟
⎜0 e
2A
k 2 +A2
t
0 ⎟
⎝ # ⎠
−2k A
t
0 0 e k 2 +A2
42 P. Baird
where λ is an arbitrary parameter. In the case when λ = 1 this gives the metric
for the geometry Sol, in particular, Sol is one of a 1-parameter family of soliton
metrics.
Example 3.2.5 (Warped product examples) We consider the special case when
a ≡ b, which can be considered as a warped product of the real line with
standard Euclidean 2-space. Now our system can be viewed as the sub-system
of the dynamical system (3.6) given by the orbits passing through the axis
u = v = 0 when A + k = 0. It is easily seen that the plane u = v is an invariant
subspace, so the orbits remain in this plane. This then leads to the 2-dimensional
Explicit constructions of Ricci solitons 43
dynamical system:
u = uρ + A
ρ = 2u2 − A .
By eliminating ρ, we see that u is given as a solution to the second order ODE:
u u − (u )2 + Au + Au2 − 2u4 = 0.
A particular solution when A = 0 is given by u = ± √12 t −1 , which yields
√
the metric g = dt 2 + t ± 2 ((dx 2 )2 + (dx 3 )2 ). This singular metric of negative
scalar curvature was also noticed in [2]. It corresponds to a steady gradient 3-
dimensional soliton. If we don’t insist on completeness, this example provides
an answer to Problem 1.88 of [5]; by the expressions for the components of the
Ricci curvature, we see that the Ricci tensor is positive whatever sign is taken
in the expression for u, as is required – cf. the discussion in [5].
Example 3.2.6 (Non-standard structure over S 2 ) Let us make the orthogonal
substitution:
Au + kv −ku + Av
ξ=√ , η= √
k +A
2 2 k 2 + A2
which puts (3.6) into the form:
⎧
⎨ ξ = ξρ √
η = ηρ + k 2 + A2 (3.9)
⎩
ρ = ξ 2 + η2 − A .
We study the orbit with initial condition (ξ (0), η(0), ρ(0)) = (0, 0, 0), so that
ξ (t) = 0 for all t, and we now have the reduced system:
√
η = ηρ + k 2 + A2
(3.10)
ρ = η2 − A .
% !
√ k 2 + A2
If A > 0, the fixed points are given by ± A, ∓ , otherwise there
A
are none unless A = k = 0 in which case the whole axis η = 0 consists of fixed
points. We study the case A = 0, k = 0; for the other cases, see [1].
The dynamical system (3.10) now takes the form
η = ηρ + k
(3.11)
ρ = η2 .
On replacing η with −η, it is no loss of generality to suppose that k > 0. We
consider the orbit with initial condition (η(0), ρ(0)) = (0, 0). Then there exists
a T > 0 such that η(t), ρ(t) → ∞ as t → T − . In [1] it is shown that this orbit
44 P. Baird
E = X + f U,
.
dθ + μ ∧ θ =
We are now ready to state an improved version of the ansatz given in [2].
46 P. Baird
where A and B are constants and where equation (iii) is vacuous if ζ ≡ 0. Set
M = N × (−δ, δ)√for some δ > 0, and let ϕ : M → N be the canonical pro-
= ϕ ∗ , ρ = ρ ◦ ϕ. Let θ be a 1-form satisfying
jection. Let = 2 ζ μN ,
,
dθ + d ln ρ ∧ θ = (3.14)
Proof : By [2], it is sufficient to prove the equivalence of the above system with
the following set of equations:
⎧
⎪
⎪
2
(i) (a) K N + 12 N ln(λ ν) − |grad ln ρ|2 + A−ψ =0
⎪
⎪ 2
⎪
⎨ (i) (b)
λ
∇
(
2
ln ν + 2d ln λ d ln ν − (d ln ρ) 2
= αh )(some α : N → R)
2
⎪
⎪ (ii) λ ln ρ − h(grad ln ρ, grad ln ν) + 2 = const.
N ψ+A
⎪
⎪ λ
⎪
⎩ (iii) −1/2
N ln(ρ 2 νψ ) + |grad ln ρ|2 − 1 |grad ln ψ|2
4
+ 12 h(grad ln ψ, grad ln ν) + ψ+A
2 = 0, (3.15)
λ
2K N + δ ln ν − |grad ln ρ|2 + 2A − 2ζ
which is equivalent to (i) (a). Thus (i) implies (i) (a) and (b). Conversly, given
(i) (a) and (b), taking the trace of (i) (b) gives
2α = ln ν − |grad ln ρ|2 ,
Explicit constructions of Ricci solitons 47
2
which by (i) (a) equals −2K N − 2A + 2ζ , so that
2
∇ 2 ln ν − (d ln ρ)2 + K N h + AH − ζ h = 0,
which is precisely (i).
Example 3.3.2 (Case of integrable horizontal distribution) If ζ ≡ 0, then the
horizontal distribution of ϕ is integrable and equation (3.13)(iii) becomes vac-
uous (cf. [2]). We then have the following coupled system to solve:
(i) Ricci N + ∇ 2 ln ν − (d ln ρ)2 + Ah = 0
(ii) ln ρ − h(grad ln ρ, grad ln ν) + B = 0 .
For example, the dynamical system of Theorem 3.2.1 can be seen to arise this
way.
Example 3.3.3 (Case of minimal fibres–the geometry Nil) The fibres of ϕ
are minimal if and only if ρ = const. In this case equation (ii) implies that
2
ζ = −B is constant and (i) becomes:
Ricci N + ∇ 2 ln ν + (A − B)h = 0 ,
which is the equation for a gradient soliton on the surface (N 2 , h). If ζ = 0,
then the horizontal distribution is integrable and we locally have a product
structure M 3 = × R, where is a 2-dimensional gradient soliton.
If on the other hand ζ = 0, then equation (iii) implies that
2
ln ν + ζ + A = 0 ,
which combined with the trace of (i) gives the identity:
2
2K N + A − 3ζ = 0 ,
so that in particular the curvature K N must be constant. If we√take N = R2 with
its canonical metric and coordinates (x, y) and set ζ = 1/ 2, then, applying
the construction of Theorem 3.3.1, we obtain = dx ∧ dy. To find θ we are
required to solve
dθ = dx ∧ dy .
so that, up to a diffeomorphism, we can take θ = xdy + dz (recalling that θ
must not vanish on vertical vectors). This then gives the soliton metric:
g = dx 2 + dy 2 + (xdy + dz)2 ,
which is the metric of the geometry Nil. It is easy to calculate the soliton flow,
which is given by
∂ ∂ ∂
E = −x −y − 2z .
∂x ∂y ∂z
This is not a gradient with respect to the metric g [2].
48 P. Baird
where {ei } is an orthonormal basis of the horizontal space (ker dϕ)⊥ and
g H = ϕ ∗ h/λ2 is the horizontal part of the metric.
Proof : As before, let U be a unit vector field tangent to the fibres of ϕ and let
θ = g(U, · ). If one rescales: V = λn−2 U , = d
θ = λ−n+2 θ , θ , then in [3],
the Ricci curvature is computed on components as follows; here we suppose
that X, Y are orthogonal to U :
We are now able to obtain the soliton equations in terms of the parameters
of a semi-conformal submersion as expressed by the following theorem.
Now suppose that t denotes a unit speed parameter along the fibres, so that
U = ∂/∂t and θ = dt, then a necessary consequence of the equations is that
f = f (t). There are product solutions given by λ =const. If, on the other hand,
we suppose that λ is non-constant and work on a neighbourhood where λ = 0,
50 P. Baird
3x1 2 x1 2
1
22 = −x1 , 1
23 = − 12 1 + 2
1
, 24 = − 12 , 1
33 = −x1 1 + 2
,
2
1
34 = − x21 , 2
13 = 1
2
1− x1
2
, 2
14 = 12 , 3
12 = 12 ,
x1 2
3
13 = x1
2
, 4
12 = 1
2
1− 2
, 4
14 = − x21 ,
1 1 x1 2 2
Ricci = −dx1 2 − dx3 2 + x1 dx2 + dx3 + dx4 .
2 2 2
We now introduce the soliton flow E = α∂1 + β∂2 + γ ∂3 + f ∂4 , compute
LE g and substitute into the soliton equation (3.2). We omit the details of this
latter calculation. When we equate the different coefficients to zero, we obtain
the following set of equations to solve:
dx1 2 : −2 + 2∂1 α + 2A = 0
x1 2
dx1 dx2 : 2∂2 α + 2(1 + x1 2 )∂1 β + 2x1 1 + 2
∂1 γ + 2x1 ∂1 f = 0
2
x1 2 x1 2
dx1 dx3 : 2∂3 α + 2x1 1 + 2
∂1 β + 2 1 + 2
∂1 γ + x1 2 ∂1 f = 0
dx1 dx4 : 2∂4 α + 2x1 ∂1 β + x1 2 ∂1 γ + 2∂1 f = 0
x1 2
dx2 2 : x1 2 + 2x1 α + 2(1 + x1 2 )∂2 β + 2x1 1 + 2
∂2 γ
+ 2x1 ∂2 f + 2(1 + x1 2 )A = 0
x1 2
dx2 dx3 : x13 + (2 + 3x1 2 )α + 2(1 + x1 2 )∂3 β + 2x1 1 + 2
∂3 γ + 2x1 ∂3 f
2 2 2
+ 2x1 1+ x21 ∂2 β+2 1+ x21 ∂2 γ +x1 2 ∂2 f + 2A(2x1 +x1 3 ) = 0
x1 2
dx2 dx4 : 2x1 + 2α + 2(1 + x1 2 )∂4 β + 2x1 1 + 2
∂4 γ + 2x1 ∂4 f
+ 2x1 ∂2 β + x1 2 ∂2 γ + 2∂2 f + 4x1 A = 0
x1 4 x1 2 x1 2
dx3 2 : −1 + 4
+ 2x1 1 + 2
α + 2x1 1 + 2
∂3 β
2 2
x1 2 x1 2
+2 1 + 2
∂3 γ + x1 2 ∂3 f + 2A 1 + 2
=0
x1 2 x1 2
dx3 dx4 : x1 2 + 2x1 α + 2x1 1 + 2
∂4 β + 2 1 + 2
∂4 γ + x1 2 ∂4 f
+ 2x1 ∂3 β + x1 2 ∂3 γ + 2∂3 f + 2Ax1 2 = 0
dx4 2 : 1 + 2x1 ∂4 β + x1 2 ∂4 γ + 2∂4 f + 2A = 0 .
52 P. Baird
ax1 2
(a) : ∂111 q = 0 ⇒ r = 0 ⇒ r = + bx1 + c ,
2
for some (new) constants a, b, c. We conclude that
p=e
2
q = −3x1 x2 − x1 2 x3 − 2ex2 − 2ex1 x3 + ax21 + bx1 + c
α = − x21 + e
ax 2
v = −4x4 − 3x1 x2 − x1 2 x3 − 2ex2 − 2ex1 x3 + 2[ + bx1 + c
w = 12 (−3x2 − 2x1 x3 − 2ex3 + ax1 + b)
u = x1 (−3x2 − 2x1 x3 − 2ex3 + ax1 + b) − 2x3 + a − 2x1 2 x4
2 ax 2
+ x21 (−3x1 x2 − x1 2 x3 − 2ex2 − 2ex1 x3 + 2[ + bx1 + c) ,
3x2 x1
β=− + a − 2ex3 + b
2 2
a
γ = −x3 +
2
f = −2x4 − ex2 .
Theorem 3.5.1 The geometry Nil4 admits a unique soliton structure, with flow
defined up to addition of a Killing vector field, given by
x1 3x2
E=− ∂1 − ∂2 − x3 ∂3 − 2x4 ∂4 .
2 2
This result is confirmed by Theorem 3.4.2, where we take ϕ : Nil4 → Nil3
to be the natural projection and exploit the soliton structure on Nil3 given
by Example 3.3.3; indeed, we apply Theorem 3.4.2 with λ ≡ 1, θ = dx4 +
2
x1 dx2 + x21 dx3 and X = − x21 ∂1 − 3x22 ∂2 − x3 ∂3 . However, 3.4.2 does not
establish uniqueness.
Explicit constructions of Ricci solitons 55
References
[1] P. Baird, A class of three-dimensional Ricci solitons, Geometry and Topology 13
(2009), 979–1015.
[2] P. Baird and L. Danielo, Three-dimensional Ricci solitons which project to surfaces,
J. reine angew. Math., 608 (2007), 65–91.
[3] P. Baird and J. C. Wood, Harmonic Morphisms between Riemannian Manifolds,
London Math. Soc. Monograph (New Series), vol. 29, Oxford University Press,
2003.
[4] R. L. Bryant, Ricci flow solitons in dimension three with SO(3)-symmetries,
preprint, Duke Univ., Jan 2005.
[5] B. Chow, S-C. Chu, D. Glickenstein, C. Guenther, J. Isenberg, T. Ivey, D. Knopf,
P. Lu, F. Luo and L. Ni, The Ricci flow: Techniques and Applications, Part 1:
Geometric aspects, AMS Mathematical Surveys and monographs, 135, 2007.
[6] B. Chow and D. Knopf, The Ricci flow: An Introduction, Mathematical Surveys
and Monographs, Vol. 110, American Mathematical Society, Providence, RI, 2004.
[7] C. Guenther, J. Isenberg and D. Knopf, Stability of the Ricci flow at Ricci-flat
metrics, Comm. Anal. Geom. 10 (2002), no. 4, 741–777.
[8] C. Guenther, J. Isenberg and D. Knopf, Stability of Ricci nilsolitons, preprint
(2006).
[9] R. Hamilton, A compactness property for solutions of the Ricci flow, Amer. J.
Math. 117 (1995), 545–572.
[10] T. Ivey, Ricci solitons on compact three-manifolds, Differential Geom Appl. 3 (4)
(1993), 301–307.
[11] T. Ivey, New examples of complete Ricci solitons, Proc. Amer. Math. Soc. 122
(1994), 241–245.
[12] J. Lauret, Ricci soliton homogeneous nilmanifolds, Math. Ann. 319 (2001), 715–
733.
[13] J. Lott, On the long-time behaviour of type-III Ricci flow solutions, Math. Annalen,
339, No. 3 (2007), 627–666.
[14] G. Perelman, The entropy formula for the Ricci flow and its geometric applications,
arXiv:math.DG/0211159.
[15] N. Sesum, Linear and dynamical stability of Ricci flat metrics, Duke. Math. J.,
133 (2006), 1–26.
Author’s address:
Département de Mathématiques,
Université de Bretagne Occidentale,
6 Avenue Le Gorgeu, – CS 93837
29238 Brest,
France
[email protected]
4
Open Iwasawa cells and applications
to surface theory
josef f. dorfmeister
4.1 Introduction
In recent years, many surfaces of a special type, like surfaces of constant mean
curvature (CMC) in R3 , Willmore surfaces in S n or spacelike mean curvature
surfaces in Minkowski space L3 , have been constructed using loop groups.
The procedure in all these cases is fairly similar: one considers a ‘Gauss type
map’ from a Riemann surface M to some real symmetric space G/K, like
the classical Gauss map or a conformal Gauss map, and characterizes a class
of surfaces by the harmonicity (or the conformal harmonicity) of this Gauss
map. Lifting the Gauss map to a map F from the universal cover M̃ into G
we obtain a ‘moving frame’. It has been shown in [7] how all such frames
can be constructed from holomorphic data: considering the moving frame for
each member of the associated family one obtains an ‘extended frame’ Fλ and
one can show that it suffices to construct all such extended frames, since, as an
added feature, for all these special surface classes there exists a simple formula,
‘Sym type formula’, which reconstructs the given immersion from its extended
frame.
The heart of the loop group method for the construction of surfaces thus
is a procedure that produces all extended frames of all surfaces of the special
classes to which this method applies. This procedure involves loop group
decompositions. For simplicity, let’s consider conformal immersions like CMC
and spacelike CMC from some simply connected domain D ⊂ C into R3 and
L3 respectively. Then the extended frame F has, for all z ∈ D \ S, S a discrete
subset of D, a ‘Birkhoff decomposition’ F = F− L+ . (Here F− only contains
non-positive powers of λ and starts with I and L+ only contains non-negative
powers of λ in a Fourier expansion.) The matrix function F− maps into GC
and is meromorphic with poles at the points of S. The Maurer-Cartan form
η = F−−1 dF− of F− is a meromorphic (1, 0)− form and is called the ‘normalized
56
Open Iwasawa cells 57
potential’ for the given immersion. It is of the form η = λ−1 η−1 (z)dz and has
values in pC , where g = k + p is the Cartan decomposition of the symmetric
space G/K. As a consequence, if one wants to construct an extended frame
for a surface in one of the classes under consideration, one will start with
some normalized potential η = λ−1 η−1 (z)dz, a meromorphic (1, 0)−form with
values in pC , which has a meromorphic solution to the ODE dF− = F− η. In
the next step one decomposes F− in the form F− = F W+ , with F ∈ G. If G is
maximal compact in the complex Lie group GC , then this decomposition always
exists and is called ‘Iwasawa decomposition’. But in many cases relevant to
surface theory G is non-compact. In the finite dimensional case such a situation
has been investigated by Aomoto, Matsuki and Rossman [1], [13], [15]. In the
(infinite dimensional) loop group case similar results have been obtained by
Kellersch [10], [11]. It turns out that a decomposition of the form F− = F W+
only exists away from a singular set S. This singular set may consist of points,
but may also contain curves. The matrix function F is, where defined, the
extended frame of a harmonic map associated with surfaces of our class. Then
a ‘Sym formula’ produces the actual surface. More precisely, the Sym formula
produces a map, called here ‘weak immersion’, which is an immersion of the
desired surface type, wherever its differential has rank two.
In the construction outlined above, three types of singularities occur: firstly,
poles in the normalized potential. Secondly, singularities due to the non-global
Iwasawa splitting and, thirdly, branch points or branching curves, where the
differential of the mapping defined by the Sym formula drops rank.
The first kind of singularities can be taken care of by considering a slightly
different construction, producing holomorphic potentials (with a less trivial
Fourier expansion). The last kind of singularities is sometimes unavoidable
and will not be discussed in this note.
Singularities of the second kind, stemming from the non-globality of the
Iwasawa splitting for non-compact G, have been investigated by a few authors
only so far and for G = Sl(2, R) only, to the best of our knowledge. It turns
out [2] that the coefficients of the frame F have strong singularities along the
singular set S, while the actual weak immersion may be smooth across S with
a differential of rank lower than two.
To understand this behaviour better, we consider general symmetric spaces
+
G/K and the corresponding Iwasawa decompositions LGC σ = δ∈ LGσ · δ ·
L+ GC σ , where is a set of representatives for the obvious action of LGσ ×
+ C C
L Gσ on LGσ . These cosets will be called ‘Iwasawa cells’. Similar to what
was found in the Sl(2, R) case one also finds, for general real G, that in general
several open Iwasawa cells exist. Clearly, considering F− as above, we obtain
a well-behaved extended frame of some harmonic map as long as F− stays in
58 J. F. Dorfmeister
Theorem 4.1.1 (1) The union of all open Iwasawa cells is dense in LGC σ.
(2) There exists a line bundle L∗ over LGC
σ and a real analytic section α∗
∗ C ∗
of L such that some g ∈ LGσ is in an open Iwasawa cell only if α (g) = 0.
1 LGC C
σ = {g ∈ LG ; σ (g(λ)) = g(−λ)},
2 L Gσ = {g ∈ LGC
+ C
σ ; g has a holomorphic extension to the open unit disk},
3 L+ G
∗ σ
C
= {g ∈ LG C
σ ; g has a holomorphic extension to the open unit disk
and g(0) = I },
4 L− GC C
σ = {g ∈ LGσ ; g has a holomorphic extension to the upper
hemisphere},
5 L− C C
∗ Gσ = {g ∈ LGσ ; g has a holomorphic extension to the upper hemisphere
and g(∞) = I },
6 LGσ = {g ∈ LGC ; g(λ) ∈ G for all λ ∈ S 1 }.
Note, we identify the unit disk with the lower hemisphere of the Riemann
sphere S 2 . And holomorphicity in the upper hemisphere means that after chang-
ing λ → λ−1 one has a holomorphic quantity on the open unit disk.
Open Iwasawa cells 59
Using this notation we can state the first important decomposition theorem
(see e.g. [14],[3]).
g = g− ωg+ (4.2)
for some g∗ ∈ L∗ GC σ and ω ∈ . The set for the disjoint union above is
closely related to the Weyl group of LGCσ (see e.g. [14]).
Moreover, the group multiplication L− GC + C C
σ × L Gσ → LGσ is an analytic
C
diffeomorphism onto an open dense subset of LGσ .
Remark 4.2.2 (1) The Birkhoff decomposition for g in the ‘big (left Birkhoff)
cell’
L− GC + C − C
σ · L Gσ can be made unique if one requires g− ∈ L∗ Gσ or g+ ∈
+ C
L∗ Gσ .
(2) For the other cells one can also formulate a condition that makes the
representation unique [9],[14].
(LGσ × L+ GC C C
σ ) × LGσ → LGσ , where (g, v+ ).h = ghv+
−1
(4.3)
g = hδg+ (4.5)
Theorem 4.4.2 With the notation of the last theorem we can assume w.l.g.
that n has the form n = qt, where q is in the normalizer of T in U and t is a
homomorphism from S 1 into T . In particular, n is a representative of a Weyl
group element of the twisted loop group LGC σ.
Proof Note that g = hv+ is equivalent with τ (g)−1 g = τ (v+ )−1 v+ . The real-
ization τ (g)−1 g = τ (v+ )−1 · qt · v+ shows that g is as required only if τ (g)−1 g
is in the big Birkhoff cell. This happens only if qt(λ) ∈ K C . Since this matrix
62 J. F. Dorfmeister
Even if the union of all open Iwasawa cells is not equal to the non-vanishing
set of the real analytic section discussed above, we still obtain
Proposition 4.5.4 The union of all open Iwasawa cells is dense in LGC
σ.
Proof If the union of all open Iwasawa cells is not dense, then there exists an
open set in GC ∗
− σ of the complement of this union such that the section α does
not vanish anywhere on this open set. To each such point there corresponds
some q as in Proposition 4.2.
64 J. F. Dorfmeister
We note that every such q can be written in the form q = aρ, where a ∈ T
and ρ is a once and for all fixed representative of the Weyl group of K C relative
to T . Thus freedom can only come from the fact that some q = aρ, but also
some q̂ = âρ occurs. If the claim of the proposition is wrong, then there exists
some q to which uncountably many q̂ s belong. But then one of the a’s would
be a cluster point. Now (2) in the corollary above shows that this is not possible
for q s representing different Iwasawa cells.
Remark 4.5.5 (1) In the finite dimensional case, Aomoto [1] has given a very
satisfactory description of all occurring cells. A fairly complete description of
all Iwasawa cells has been given in the untwisted case by Kellersch [10],[11].
For the twisted case much remains to be done.
need to consider conformally harmonic maps from D into the symmetric space
Gr1,3 (Ln+2 ) ∼
= SO(1, n + 1)/SO(1, 3) × SO(n − 2). As a consequence, the
loop group method is applicable.
In this case the complex Lie group GC is SO(n + 2, C) with real form
∼
G = SO(1, n + 1), given by the involution τ (X) = T0 X̄T0 , where T0 is the
diagonal matrix with diagonal entries −1, 1, 1, . . . and the symmetric space
structure is given by the involution σ (B) = SBS, where S is a diagonal matrix,
with the first four entries equal to −1 and all other entries equal to 1.
Thus K C = SO(4, C) × SO(n − 2, C). The general theory we have pre-
sented in this paper does allow for many open Iwasawa cells and we believe
that they actually exist and will give rise to many effects for general Willmore
surfaces in S n . (We have not yet computed all Iwasawa cells in this example.)
However, fortunately, for the case of Willmore spheres things are very
simple. It turns out [8] that in this case the corresponding conformal harmonic
maps are of ‘finite uniton number’ and therefore the corresponding normalized
potentials can be represented as upper triangular matrices. As a consequence,
the Iwasawa splittings can be carried out explicitly and it seems that in this
case of Willmore spheres the matrix function C(z, λ) always stays in the open
Iwasawa cells which contains I . Hence the problems discussed in Example 1
will not occur for Willmore spheres (but possibly for Willmore tori).
References
[1] K. Aomoto, On some double coset decompositions of complex semisimple Lie
groups, J.Math.Soc.Japan 18 (1966), 1–44
[2] D. Brander, W. Rossman, N. Schmitt, Holomorphic representation of constant
mean curvature surfaces in Minkowski space:consequences of non-compactness
in loop group methods, Adv.Math. 223 (2010), 949–986
[3] J. Dorfmeister, H. Gradl, J. Szmigielski, Systems of PDEs obtained from factor-
ization in loop groups, Acta Appl.Math. 53 (1998), 1–58
Open Iwasawa cells 67
Authors’ addresses:
Josef F. Dorfmeister
Fakultät für Mathematik,
Technische Universität München,
Boltzmannstr. 3,
D-85747 Garching,
Germany
[email protected]
5
Multiplier ideal sheaves and
geometric problems
akito futaki and yuji sano
Abstract
In this expository article we first give an overview on multiplier ideal sheaves
and geometric problems in Kählerian and Sasakian geometries. Then we review
our recent results on the relationship between the support of the subschemes cut
out by multiplier ideal sheaves and the invariant whose non-vanishing obstructs
the existence of Kähler-Einstein metrics on Fano manifolds.
5.1 Introduction
One of the main problems in Kählerian and Sasakian geometries is the exis-
tence problem of Einstein metrics. An obvious necessary condition for the
existence of a Kähler-Einstein metric on a compact Kähler manifold M is that
the first Chern class c1 (M) is negative, zero or positive since the Ricci form
represents the first Chern class. This existence problem in Kählerian geometry
was settled by Aubin [1] and Yau [58] in the negative case and by Yau [58]
in the zero case. In the remaining case when the manifold has positive first
Chern class, in which case the manifold is called a Fano manifold in algebraic
geometry, there are two known obstructions. One is due to Matsushima [29]
which says that the Lie algebra h(M) of all holomorphic vector fields on a
compact Kähler-Einstein manifold M is reductive. The other one is due to
the first author [16] which is given by a Lie algebra character F : h(M) → C
with the property that if M admits a Kähler-Einstein metric then F vanishes
identically. Besides, it has been conjectured by Yau [59] that a more subtle con-
dition related to geometric invariant theory (GIT) should be equivalent to the
existence of Kähler-Einstein metrics. This idea was made explicit in the paper
[50] of Tian in which a notion called K-stability was introduced. Tian used a
68
Multiplier ideal sheaves 69
generalized version of the invariant F for normal almost Fano varieties and
used it as the numerical invariant for the stability condition. The link between
the idea of GIT stability and geometric problems such as the existence problems
of Hermitian-Einstein metrics and constant scalar curvature Kähler metrics can
be explained through the moment maps in symplectic geometry. The explana-
tion from this viewpoint can be found for example in [14], [15], [13]. Recall
that an extremal Kähler metric is by definition a Kähler metric such that the
gradient vector field of the scalar curvature is a holomorphic vector field. In
particular, a Kähler metric of constant scalar curvature is an extremal Kähler
metric. The theorem of Matsushima is extended for extremal Kähler manifolds
by Calabi [3] as a structure theorem of the Lie algebra h(M) on an extremal
Kähler manifold M, and the first author’s obstruction F can be extended as
an obstruction to the existence of constant scalar curvature Kähler metric in a
fixed Kähler class ([17], [3]). The theorem of Calabi and the character F are
explained in the framework of the moment map picture by X. Wang [55] (see
also [19]).
In [10] Donaldson refined the notion of K-stability for a polarized manifold
(M, L), that is, a pair of an algebraic manifold M and an ample line bundle
L over M, and conjectured that there would exist a Kähler form in c1 (L)
of constant scalar curvature if and only if (M, L) is K-polystable. To define
K-(poly)stability for (M, L) Donaldson refined the invariant F even for non-
normal varieties which are degenerations of the polarized manifold (M, L) and
used it as the numerical invariant for the stability condition. The notion of
K-stability is defined as follows. For an ample line bundle L over a projective
variety M of dimension m, a test configuration of exponent r consists of the
following:
lar for r sufficiently large dim H 0 (Mt , Lrt ) = dim H 0 (M, Lr ) for all t ∈ C.
Here we write Lrt for L|Mt though L may not exist for t = 0.
2
ξ to S ∼
Then (ξ − iJ ξ ) is a holomorphic vector field. The restriction ξ of
1
=
{r = 1} becomes a Killing vector field, called the Reeb vector field. The flow
generated by ξ is called the Reeb flow. The restriction η, of the 1-form η on
C(S) defined as
1 √
η = 2 ḡ(ξ̃ , ·) = −1(∂¯ − ∂) ln r,
r
∼
to S = {r = 1} becomes a contact form. Hence dη defines Kähler forms on
local orbit spaces of Reeb flow. That is to say, the 1-dimensional foliation
defined by ξ comes equipped with a structure of transverse Kähler foliation.
A Sasaki manifold is said to be regular if the Reeb flow generates a free S 1 -
action, quasi-regular if all the orbits are closed. A Sasaki manifold is said to be
irregular if it is not quasi-regular.
For a polarized manifold (M, ω) the associated U (1)-bundle S of L becomes
a regular Sasaki manifold in a natural way: Choose a positive (1, 1) form
representing c1 (L), take the Hermitian metric h on L such that the connection
form η on L has its curvature form d η equal to ω. The Kähler cone C(S) is
L minus the zero section with the Kähler form given by 2i ∂∂r 2 where r is
the distance from the zero section. Conversely, any regular Sasaki manifold
is given in this way. Similarly, a quasi-regular Sasaki manifold is given as an
associated U (1)-orbibundle over an orbifold.
As is shown in [20] most of ideas in Kähler geometry can be extended to
transverse Kähler geometry for Sasaki manifolds. For example one can extend
Calabi’s theorem to compact Sasaki manifold whose transverse Kähler metric
is an extremal Kähler metric, and one can extend the obstruction F as an
obstruction for a basic cohomology class to admit a transverse Kähler form
with constant scalar curvature.
A Sasaki-Einstein manifold is a Sasaki manifold whose metric is an Ein-
stein metric. This condition is equivalent to the transverse Kähler metric being
Kähler-Einstein. Thus the study of the existence problem of Sasaki–Einstein
metrics are closely related to the problem of Kähler–Einstein metrics. But there
are differences between them. To explain the differences let k be the maximal
dimension of the torus which acts on C(S) by holomorphic isometries. When
k = m + 1 the cone C(S) is a toric variety, and in this case the Sasaki manifold
S is said to be toric. Notice that k is at least 1 because
ξ generates holomorphic
72 A. Futaki and Y. Sano
for any γ ∈ (γ0 , 1), and that there is a nonempty open subset U ⊂ M satisfying
e−ϕk dV ≤ O(1) (5.3)
U
as k → ∞, where dV is a fixed volume form. Note that the last condition (5.3)
always holds for any S due to gi j¯ + ∂i ∂j¯ ϕ > 0 (see for instance [49]). For each
S, Nadel constructed a coherent ideal sheaf I(S), which is called the multiplier
ideal sheaf (MIS). We will explain later the simpler definition of multiplier
ideal sheaves given by Demailly–Kollár [9].
Here let us recall the outline of Nadel’s construction. (See the original paper
[30] for the full details.) Let L be an arbitrary ample line bundle on M which
is not necessarily the anticanonical line bundle of M. We define H 0 (M, Lν )S
to be the set of all f ∈ H 0 (M, Lν ) for which there exists a sequence {fk } of
H 0 (M, Lν ) such that
|fk |2 e−γ ϕk dV ≤ C
M
Define
∞
-
I (M, S, L) = H 0 (M, Lν )S ,
ν=0
which is a homogeneous ideal I (M, S, L) of the graded ring R(M, L). Then,
the ideal sheaf I(S) is defined as the algebraic sheaf of ideals on M associated
to I (M, S, L). It is proved in [30] that this construction is independent of the
choice of L. Let V(S) be the (possibly non-reduced) subscheme in M cut out by
I(S). This subscheme is characterized as follows. A point p ∈ M is contained
in the complement of V(S) if and only if there exist an open neighborhood W
of p in M and a real number γ ∈ (γ0 , 1) such that
e−γ ϕk dV ≤ O(1)
W
74 A. Futaki and Y. Sano
√
−1 ¯
Ric(g) − ωg = ∂ ∂hg , ehg ωgm = ωgm = V . (5.7)
2π M M
(5.6) violates the above estimate, then there is a sequence {tk } such that tk → t0
as k → ∞ and {ϕtk − supM ϕtk }∞ k=1 induces a proper multiplier ideal sheaf I.
In this paper, we call it the Kähler-Einstein multiplier ideal sheaf (KE-MIS).
Summing up,
Theorem 5.2.2 ([30]) Let M be a Fano manifold which does not admit a
Kähler-Einstein metric. Let G be a compact subgroup of the group Aut(M)
of holomorphic automorphisms of M. Assume that M does not admit any
GC -invariant proper multiplier ideal sheaf. Then M admits Kähler-Einstein
metrics. Here GC denotes the complexification of G.
By combining the above theorem and the geometric properties of V(S) given
by the vanishing formula (5.5), Nadel gave many examples of Kähler-Einstein
Fano manifolds. Recently Heier [23] applied this method to (re-)prove the
existence of Kähler-Einstein metrics on complex del Pezzo surfaces obtained
from the blow up of CP2 at 3,4 or 5 points, which was originally proved by Siu
[44], Tian [49], Tian and Yau [51].
This method was extended to the case of Fano orbifolds by Demailly-Kollár
[9]. Their construction is simpler than [30]. Let ψ be an ωg -plurisubharmonic
(psh) function (or almost psh function with respect √
to ωg ), i.e., a real-valued
−1 ¯
upper semi-continuous function satisfying ωg + 2π ∂ ∂ψ ≥ 0 in the current
sense. The multiplier ideal sheaf with respect to ψ in the sense of [9] is the
ideal sheaf defined by the following presheaf
(U, I(ψ)) = {f ∈ O(U ) | |f |2 e−ψ dV < ∞} (5.8)
U
where U is an open subset of M. This sheaf is also coherent and satisfies the
vanishing theorem of Nadel type. In terms of this formulation, Theorem 5.2.2
can be written as follows. Let {ϕt } be the solution {ϕt }0≤t<t0 of (5.6) which
violates a priori estimates.
r tk → t0 as k → ∞,
r there exists a limit ϕ∞ = limk→∞ (ϕtk − supM ϕtk ) in the L1 -topology, which
is an ωg -psh function, and
r I(γ ϕ∞ ) is a GC -invariant proper multiplier ideal sheaf, i.e, I(γ ϕ∞ ) is
neither 0 nor OM .
76 A. Futaki and Y. Sano
if γ < 1. This means that the positivity condition for (5.9) with respect to
h0 e−γ ϕ holds if γ < 1. Then (5.9) implies
Lemma 5.2.4 If there exists a positive constant γ < 1 and an ωg -psh function
ϕ such that I(γ ϕ) is proper, then I(γ ϕ) satisfies (5.4). In particular, I(γ ϕ∞ )
for γ ∈ (m/(m + 1), 1) in Theorem 5.2.3 satisfies (5.4) (and then (5.5)).
On the other hand, the lower bound of γ in Theorem 5.2.3 describes the
strength of the singularity of ϕ∞ . It is closely related to a holomorphic invariant
introduced by Tian [49]. It is often called the α-invariant, which is defined by
αG (M) := sup{α ∈ R | e−α(ψ−supM ψ) ωgm < Cα for all G-invariant ωg -pshψ}
M
(5.10)
where G ⊂ Aut(M) is a compact subgroup. If a multiplier ideal sheaf I(γ ψ)
with respect to a G-invariant ωg -psh function ψ of exponent γ is proper,
where supM ψ = 0, then αG (M) ≤ γ , because e−γ ψ is not integrable over M.
Conversely,
Multiplier ideal sheaves 77
Lemma 5.2.5 If αG (M) < 1, then there exist a positive constant γ ∈ (0, 1)
and a G-invariant ωg -psh function ψ with supM ψ = 0 such that I(γ ψ) is
proper.
Tian gave a sufficient condition for the existence of Kähler-Einstein metrics
on Fano manifolds in terms of this invariant.
Theorem 5.2.6 ([49]) If αG (M) > m/(m + 1), then M admits a G-invariant
Kähler-Einstein metric.
Using Theorem 5.2.6, Tian and Yau [51] proved the existence problem of
Kähler-Einstein metrics on Fano surfaces, i.e., the Fano surfaces obtained from
the blow up of CP2 at k points where 3 ≤ k ≤ 8 admits a Kähler-Einstein metric.
Both of the lower bound of αG (M) and the non-existence of the proper multiplier
ideal sheaves satisfying (5.5) give sufficient condition for the existence of
Kähler-Einstein metrics on Fano manifolds, and they are related directly to
each other. For example, Lemma 5.2.4 and 5.2.5, we have
Lemma 5.2.7 If αG (M) < 1, then a GC -invariant proper multiplier ideal
sheaves satisfying (5.5) exists.
Although αG (M) is difficult to compute in general, it is possible to calculate
it when M has a large symmetry in such cases as [46] for toric varieties and [12]
for the Mukai-Umemura 3-folds. On the other hand, there is a local version of
the αG (M)-invariant, which is called the complex singularity exponent [9].
Let K ⊂ M be a compact subset and ψ be a G-invariant ωg -psh function on
M. Then the complex singularity exponent cK (ψ) of ψ with respect to K is
defined by
cK (ψ) = sup{c ≥ 0 | e−cψ is L1 on a neighborhood of K}.
This constant depends only on the singularity of ψ near K. It is obvious
that cK (ψ) ≥ αG (M). One of the important properties of cK (ψ) is the semi-
continuity with respect ψ. Let P(M) be the set of all locally L1 ωg -psh functions
on M with L1 -topology. Then, we have (cf. Effective version of Main Theorem
0.2 in [9])
Theorem 5.2.8 ([9]) Let K ⊂ M be a compact subset of M. Let ϕ ∈ P(M)
be given. If c < cK (ϕ) and ψj → ϕ in P(M) as j → ∞, then e−cψj → e−cϕ
in L1 -norm over some neighborhood U of K.
In particular, if {ψj } satisfies
e−γ ψj dV → ∞
M
78 A. Futaki and Y. Sano
for any γ ∈ (m/(m + 1), 1). In Theorem 5.2.2, a subsequence of {ϕt } induces
the KE-MIS, which is proper. On the other hand, Theorem 5.2.8 implies that
e−γ ϕ∞ is not integrable over M for any γ ∈ (m/(m + 1), 1), where ϕ∞ :=
limi→∞ (ϕti − sup ϕti ). This means that ϕ∞ induces the KE-MIS in Theorem
5.2.3.
The multiplier ideal sheaves in [9] and the complex singularity exponent
can be defined algebraically as follows (cf. [27] and [2] for instance). Here we
consider a smooth variety M of dimension m. Let D = ai Di be a Q-divisor
on M. A log resolution of (M, D) is a projective birational map μ : M → M
with M smooth such that the divisor
μ∗ D + Ei
i
has simple normal crossing support. Assume D is effective and fix a log
resolution μ of (M, D). Then the multiplier ideal sheaf I(M, D) ⊂ OM with
respect to D is defined by
where !KM + D" means the integral part of KM + D. Note that I(M, D) is
independent of the choice of μ. This (algebraic) ideal sheaf corresponds to
the following (analytic) multiplier ideal sheaf defined in [9]. Take an open set
U ⊂ M so that for each Di there is a holomorphic function gi locally defining
Di in U . Let ϕD := i 2ai ln |gi | which is plurisubharmonic on U and define
|f |2
(U, I(ϕD )) := {f ∈ OM (U ) | . ∈ L1loc } (5.12)
|gi |2ai
as before. For simplicity, we assume that D = i ai Di has simple normal
crossing support. The holomorphic function f satisfies the L2 -integrability
.
condition in (5.12) if and only if f can be divided by g mi where mi ≥ !ai ",
i.e., I(ϕD ) = OM (−!D"). Let μ : M → M be a log resolution of D. Then we
have
The second equality in the above was proved in Lemma 9.2.19 [27]. We also
have
Note that this holds for an (M, D) where D does not necessarily have simple
normal crossing support (Proposition 3.20 [24]). By using the KLT condition
(5.14), we can rephrase Theorem 5.2.3. Assume that a Fano manifold M does
not admit a Kähler-Einstein metric. Let ϕt be the solution of (5.6) where
t ∈ [0, t0 ). As explained before, by taking a subsequence of {ϕtj − supM ϕtj },
there exists a limit ϕ∞ in L1 -topology, which is an ωg -psh function, such
that e−γ ϕ∞ ∈ L1 for all γ ∈ (m/(m + 1), 1). Since an approximation theorem
in [9] implies that any ωg -psh function can be approximated by an ωg -psh
function formed of ln( i |fi |2 ) where all fi are holomorphic functions, we
can replace the above ϕ∞ by an ωg -psh function formed of 2s ln |τs | where
−s
τs ∈ H 0 (M, KM ) for sufficiently large s. That is to say, there exist a sufficiently
−s 2γ
) such that e−2γ s ln |τs | = |τs |− s ∈ L1 for
1
large integer s and τs ∈ H 0 (M, KM
−s
all γ ∈ (m/(m + 1), 1). Here | · | is the induced Hermitian metric on KM with
respect to the Kähler metric g. Hence we have the following theorem. Note that
the original result holds for orbifolds, but for simplicity we assume that M is
smooth in this paper.
lctZ (M, D) := sup{λ ∈ Q | the pair (M, λD) is log canonical along Z}.
Here, the pair (M, D) is called log canonical along Z if I(M, (1 − ε)D) is trivial
in a neighborhood of every point x ∈ Z for all 0 < ε < 1. For instance, let us
−l
consider a simple case. Let M be a Fano manifold and σ ∈ H 0 (M, KM ). Let ψσ
be an ω-psh function defined by ψσ (z) = 1l ln |σ (z)|. Let Dσ be the associated
divisor with σ and Z be a closed subvariety in M. In this case, lctZ (M, 1l D) is
the same as cZ (ψσ ). The log canonical threshold plays an important role in the
studies of the multiplier ideal (sheaves) in algebraic geometry (cf. [27]). Hence
we could expect that the complex singularity exponent with respect to the limit
ϕ∞ in Theorem 5.2.3 has something to do with the existence of Kähler-Einstein
metrics although it is not clear at the moment.
To find Kähler-Einstein metrics on Fano manifolds, there is another way
instead of solving (5.6), which is the (normalized) Kähler-Ricci flow. The
Ricci flow was introduced by R. Hamilton, and on a Fano manifold M with
Kähler class c1 (M) it is defined by
d
ωt = −Ric(ωt ) + ωt , ω0 = ωg (5.15)
dt
where t ∈ [0, ∞) and ωt is the Kähler form of the evolved Kähler metric gt .
Note that (5.15) is normalized so that the Kähler class of gt is preserved. The
existence and uniqueness of the solution of (5.15) for t ∈ [0, ∞) was proved
by Cao [4]. If (5.15) converges in C ∞ , the limit is a Kähler–Einstein metric.
Then, it is natural to ask whether or not the results about the multiplier ideal
sheaves obtained from the continuity method also hold for the Kähler–Ricci
flow. The first result of this issue was given by Phong–Sesum–Sturm [37] (see
also [36]). The equation (5.15) can be reduced to the equation at the potential
level
d
ϕt = ln(ωtm /ωgm ) + ϕt − hg , ϕ0 = c (5.16)
dt
√
−1 ¯
where c is a constant and ωt = ωg + 2π ∂ ∂ϕt . They gave a necessary and
sufficient condition condition for the convergence of ϕt as t → ∞. Their proof
consists of the parabolic analogue of Yau’s arguments for the elliptic Monge-
Ampère equation, the estimates about the Kähler-Ricci flow by Perelman
(cf. [43]) and the result about the Monge-Ampère equations by Kolodziej
([25], [26]).
such that
1
sup e−pϕt ωgm < ∞.
t≥0 V M
as i → ∞ for any p > 1. Hence, the limit ϕ∞ := lim(ϕti − sup ϕti ) implies
the multiplier ideal sheaf I(pϕ∞ ), which is proper for any p > 1. That is to
say, if there is no G-invariant ωg -psh function ψ such that I(pψ) is proper for
any p ∈ (1, +∞), then M admits a G-invariant Kähler-Einstein metric. More
precisely,
Theorem 5.2.12 [37] Let M be a Fano manifold. Let G ⊂ Aut(M) be a com-
pact subgroup. Assume that M does not admit Kähler-Einstein metrics. Let
p ∈ (1, ∞) and ωg ∈ c1 (M). There is a G-invariant subsequence of the solu-
tions {ϕkj }j ≥1 of (5.16) such that
82 A. Futaki and Y. Sano
r there exists the limit ϕ∞ = limj →∞ (ϕkj − 1 ϕ ωm ) in L1 -topology,
V M kj g
which is an ωg -psh function, and
r I(pϕ∞ ) is a GC -invariant proper multiplier ideal sheaf satisfying
−!p"
H i (M, I(pϕ∞ ) ⊗ KM ) = 0, for all i ≥ 1.
Note that Nadel’s vanishing formula (5.5) need not hold for the induced MIS
I(pϕ∞ ), because p > 1. However, this result still has an application. By using
a weaker version of Nadel’s vanishing theorem and Corollary 5.2.11, Heier [23]
proved the existence of Kähler-Einstein metrics for certain del Pezzo surfaces
with large automorphism group.
After [37], Rubinstein [40] gave an analogous result as Theorem 5.2.3 for
the Kähler-Ricci flow by using a static MIS as Demailly-Kollár. His proof is
similar to the case of the continuity method, and makes use of the estimates of
Perelman and the uniform Sobolev inequality of the Kähler-Ricci flow given
by Ye [60] and Zhang [61], which appeared after [37], in stead of Kolodziej’s
theorem.
In this paper, we call the above multiplier ideal sheaf I(γ ϕ∞ ) the KRF
multiplier ideal sheaf of exponent γ . There are some remarks about the
above theorem. First, the KRF multiplier ideal sheaf is independent of the
choice of initial constant c0 of (5.16) due to the normalization of ϕti . In fact,
if we choose another constant c0 instead of the constant c0 in Theorem 5.2.13,
which is the same as Theorem 5.2.10, the solution of (5.16) is given by ϕt =
ϕt + (c0 − c0 )et . In contrast to this, when we consider the convergence of non-
normalized Kähler potentials {ϕt } as Theorem 5.2.10, we need to pay attention
to the choice of the constant c0 . Second, the normalization in Theorem 5.2.13
is equivalent to the one in Theorem 5.2.3. In fact, there is a uniform constant
C such that supM ϕt − C ≤ V1 M ϕt ωm ≤ supM ϕt . Third, γ is contained in
the interval (m/(m + 1), 1). This means that the subscheme cut out by I(γ ϕ∞ )
satisfies (5.5) and we can make use of the induced geometric properties. Fourth,
the process to prove Theorem 5.2.13 is similar to the case of the continuity
Multiplier ideal sheaves 83
m
method, and the proof in [40] implies immediately that if αG (M) > m+1 then
the Kähler-Ricci flow will converge. (This similarity is pointed out in [7] after
[40] too.)
Rubinstein [41] also gave the analogous result of Theorem 5.2.12 in terms
of the discretization of the Kähler-Ricci flow called “Ricci iteration." Given a
Kähler form ω ∈ c1 (M) and a real number τ > 0, the time τ Ricci iteration is
defined by the sequence {ωkτ }k≥0 satisfying
ωkτ = ω(k−1)τ + τ ωkτ − τ Ric(ωkτ ) for k ∈ N, (5.18)
and ω0 = ω. When τ = 1, (5.18) is the discretization of (5.15). Let Hω be the
space of Kähler potentials with respect to (ω, c1 (M))
√
∞ −1 ¯
Hω := {ψ ∈ CR | ωψ = ω + ∂ ∂ψ > 0}.
2π
Let hωψ be the Ricci potential with respect to ωψ defined as (5.7). Since
[ωkτ ] = [ω(k−1)τ ], so (5.18) can be written as the system of complex Monge-
Ampère equations
ωψmkτ = ωm ehω + τ ϕkτ −ψkτ = ωψm(k−1)τ e( τ −1)ϕkτ − τ ϕ(k−1)τ ,
1 1 1
(5.19)
where k ∈ N, ωψkτ = ωkτ and ϕkτ := ψkτ − ψ(k−1)τ .
Theorem 5.2.14 [41] Let M be a Fano manifold. Let G ⊂ Aut(M) be a com-
pact subgroup. Assume that M does not admit Kähler-Einstein metrics. Let
τ = 1. Let γ ∈ (1, ∞) and ω ∈ c1 (M). There is a G-invariant subsequence of
the solutions {ψkj }j ≥1 of (5.19) such that
r there exists the limit ϕ∞ = limj →∞ (ψkj − 1 ψ ωm ) in L1 -topology,
V M kj
which is an ω-psh function, and
r I(γ ϕ∞ ) is a GC -invariant proper multiplier ideal sheaf satisfying
−!γ "
H i (M, I(γ ϕ∞ ) ⊗ KM ) = 0, for all i ≥ 1.
Considering Yau’s conjecture, it is also natural to ask how stability conditions
in the sense of GIT is related to the convergence of the Kähler-Ricci flow. For
example, see [35], [48] and [54] for references on this issue.
where Zero(v) is the zero set of v and div(v) is the divergence of v with respect to
some Kähler metric g, i.e., div(v) = (Lv (ωgm ))/ωgm , Lv being the Lie derivative
along v. Note that Z + (v) does not depend on the choice of g, although div(v)
does. Since M does not admit a Kähler-Einstein metric, the closedness of the
set of t s for which the solutions {ϕt } of (5.6) exist does not hold, that is, the
solutions cease to exist at some t0 ∈ (0, 1]. Then the main result in [31] is as
follows.
Theorem 5.3.1 ([31]) Let M, h(M) and {ϕt } be as above. Let V be the induced
KE-MIS obtained from a subsequence of {ϕti }i where ti < t0 and ti → t0 . Then,
for any v ∈ h(M) with F (v) = 0, the support of V is not contained in Z + (v).
where uX (p) = μ(p), X and Ricω − ω = i∂∂h. Note that this normalization
is equivalent to requiring uX to satisfy
uX + Xh + uX = 0,
see [18]. For ξ ∈ tr put
D ≤0 (ξ ) := {y ∈ μ(M) | < y, ξ > ≤ 0}.
Theorem 5.3.2 ([21]) Suppose M does not admit a Kähler-Einstein metric,
and let V be the support of the KE-MIS. Let ξ ∈ tr ⊂ h(M) satisfy F (vξ ) > 0
where vξ is the holomorphic vector field corresponding to ξ . Then
μg (V ) ⊂ D ≤0 (ξ )
for any G-invariant Kähler metric g whose Kähler form is in c1 (M).
Corollary 5.3.3 Let M be the one-point blow-up of CP2 . Then V is the excep-
tional divisor.
Note that this V destabilizes slope stability in the sense of Ross-Thomas by
a result of Panov and Ross [32].
Here is the outline of the proof of Theorem 5.3.2. Let h ∈ C ∞ (M) satisfy
Ricg − ωg = i∂∂h. Suppose
det(gij + ϕij )
= e−tϕ+h
det(gij )
has solutions only for t ∈ [0, t0 ), t0 < 1. Then we have an MIS with support
V . The following fact is due to Nadel based on earlier estimates by Siu and
Tian.
Fact 5.3.4 Let K ⊂ M − V be a compact subset of M − V . Then
ωgmt → 0
K
as t → t0 .
Fact 5.3.5
μg (p) ∈ D ≤0 (ξ ) ⇐⇒ (div(vξ ))(p) ≥ 0
where
div(vξ )(eh ωm ) = Lvξ (eh ωm ).
Fact 5.3.6
t
F (vξ ) = div(vξ )ωtm .
t −1 M
Multiplier ideal sheaves 87
By Fact 5.3.6 and our assumption F (vξ ) > 0, we have for t ∈ (δ, t0 ) with
t0 < 1
t
div(vξ ) ωtm = F (vξ ) < −C
M t − 1
with C > 0 independent of t.
We seek a contradiction by assuming μg (V ) ⊂ D ≤0 (ξ ) = {div(vξ ) ≥ 0}.
Choose ! > 0 small and put
as t → t0 .
But then
−C ≥ div(vξ )ωtm = div(vξ )ωtm + div(vξ )ωtm
M M−W! W!
≥ −2!vol(M, g)
Ric(ωg ) − ωg = Lv (ωg ).
The solution for t = 1 gives the Kähler-Ricci soliton. Zhu [62] has shown that
t = 0 always has a solution. The implicit function theorem shows for some
! > 0, all t ∈ [0, !) have a solution.
Suppose we only have solutions on [0, t∞ ), t∞ < 1.
Let θv,g satisfy
iv ωg = i∂θv,g , e ωg =
θv,g m
ωgm .
M M
Tian and Zhu [52] showed that this Fv is independent of g with ωg ∈ c1 (M).
Theorem 5.3.9 (Tian-Zhu [52]) There exists a unique v ∈ hr (M) such that
Note that the assumption of NRW (M) is constrained and it would be expected
to be removed. Using the above theorem, the second author computed the
support of KRF-MIS for various γ in some examples. For example, we can
prove
Corollary 5.3.14 Let M be the blow up of CP2 at p1 and p2 . Let E1 and E2
be the exceptional divisors of the blow up, and E0 be the proper transform of
p1 p2 of the line passing through p1 and p2 . Then, the support of KRF-MIS on
M of exponent γ is
2
∪i=0 Ei for γ ∈ ( 21 , 1),
E0 for γ ∈ ( 31 , 12 ).
References
[1] T. Aubin : Equations du type de Monge-Ampère sur les variétés kählériennes
compactes, C. R. Acad. Sci. Paris, 283, 119–121 (1976)
[2] C.P. Boyer and K. Galicki : Sasakian geometry, (Oxford Mathematical Mono-
graphs., 2008).
[3] E. Calabi : Extremal Kähler metrics II, Differential geometry and complex analysis,
(I. Chavel and H.M. Farkas eds.), 95–114, Springer-Verlag, Berline-Heidelberg-
New York, (1985).
[4] H.D. Cao : Deformation of Kähler metrics to Kähler-Einstein metrics on compact
Kähler manifolds, Invent. Math. 81 (1985) 359–372.
[5] I. Cheltsov and C. Shramov : Log canonical thresholds of smooth Fano threefolds
(with an appendix by J.P. Demailly), Uspekhi Mat. Nauk 63 (2008), no. 5(383),
73–180; translation in Russian Math. Surveys 63 (2008), no. 5, 859–958.
[6] X.X. Chen and G. Tian : Geometry of Kähler metrics and foliations by holo-
morphic discs, Publ. Math. Inst. Hautes Études Sci. No. 107 (2008), 1–107.
math.DG/0409433
[7] X.X. Chen and B. Wang : Remarks on Kähler Ricci flow, arXiv:0809.3963 (2008).
[8] K. Cho, A. Futaki and H. Ono : Uniqueness and examples of toric Sasaki-Einstein
manifolds, Comm. Math. Phys., 277 (2008), 439–458, math.DG/0701122
[9] J.P. Demailly and J. Kollár : Semi-continuity of complex singularity exponents
and Kähler-Einstein metrics on Fano orbifolds, Ann. Sci. École Norm. Sup. (4)
34, no.4 (2001) 525–556.
[10] S.K. Donaldson : Scalar curvature and stability of toric varieties, J. Differential
Geometry, 62(2002), 289–349.
[11] S.K. Donaldson : Lower bounds on the Calabi functional, J. Differential Geometry,
70(2005), 453–472.
Multiplier ideal sheaves 91
[12] S.K. Donaldson : Kähler geometry on toric manifolds, and some other manifolds
with large symmetry, Handbook of geometric analysis. No. 1, 29–75, Adv. Lect.
Math. (ALM), 7, Int. Press, Somerville, MA, 2008. arXiv:0803.0985 (2008).
[13] S.K. Donaldson : Remarks on gauge theory, complex geometry and four-manifold
topology, in ‘Fields Medallists Lectures’ (Atiyah, Iagolnitzer eds.), World Scien-
tific, 1997, 384–403.
[14] S.K. Donaldson and P.B. Kronheimer : The geometry of four manifolds, Oxford
Mathematical Monographs, Claren Press, Oxford, 1990.
[15] A. Fujiki : Moduli space of polarized algebraic manifolds and Kähler metrics,
Sugaku Expositions, 5(1992), 173–191.
[16] A. Futaki : An obstruction to the existence of Einstein Kähler metrics, Invent.
Math. 73, 437–443 (1983).
[17] A. Futaki : On compact Kähler manifolds of constant scalar curvature, Proc. Japan
Acad., Ser. A, 59, 401–402 (1983).
[18] A. Futaki : Kähler-Einstein metrics and integral invariants, Lecture Notes in Math.,
vol.1314, Springer-Verlag, Berline-Heidelberg-New York, (1988).
[19] A. Futaki : Stability, integral invariants and canonical Kähler metrics, Proc. 9-th
Internat. Conf. on Differential Geometry and its Applications, 2004 Prague, (eds.
J. Bures et al), 2005, 45–58, MATFYZPRESS, Prague.
[20] A. Futaki, H. Ono and G. Wang : Transverse Kähler geometry of Sasaki manifolds
and toric Sasaki-Einstein manifolds, J. Differential Geom. 83 (2009), no. 3, 585–
635.
[21] A. Futaki and Y. Sano : Multiplier ideal sheaves and integral invariants on toric
Fano manifolds, arXiv:0711.0614 (2007).
[22] G. Heier : Convergence of the Kähler-Ricci flow and multiplier ideal sheaves on
Del Pezzo surfaces, Michigan Math. J. 58 (2009), no. 2, 423–440. arXiv:0710.5725
[23] G. Heier : Existence of Kähler-Einstein metrics and multiplier ideal sheaves on
Del Pezzo surfaces, Math. Z. 264 (2010), no. 4, 727–743. arXiv:0710.5724.
[24] J. Kollár : Singularities of pairs. (Algebraic Geometry–Santa Cruz 1995), Proc.
Sympos. Pure Math., 62, part 1 (1995) 221–287.
[25] S. Kolodziej : The complex Monge-Ampère equation, Acta Math. 180, no.1 (1998)
69–117.
[26] S. Kolodziej : The Monge-Ampère equation on compact Kähler manifolds, Indiana
Univ. Math. J. 52, no.3 (2003) 667–686.
[27] R. Lazarsfeld : Positivity in Algebraic Geometry II: Positivity for vector bundles,
and multiplier ideals, (Springer-Verlag, 2004).
[28] T. Mabuchi : K-stability of constant scalar curvature polarization, arXiv:0812.
4093
[29] Y. Matsushima : Sur la structure du groupe d’homéomorphismes d’une certaine
variété kaehlérienne, Nagoya Math. J., 11, 145-150 (1957).
[30] A.M. Nadel : Multiplier ideal sheaves and Kähler-Einstein metrics of positive
scalar curvature, Ann. of Math. (2) 132, no.3 (1990) 549–596.
[31] A.M. Nadel : Multiplier ideal sheaves and Futaki’s invariant, Geometric Theory of
Singular Phenomena in Partial Differential Equations (Cortona, 1995), Sympos.
Math., XXXVIII, Cambridge Univ. Press, Cambridge, 1998. (1995) 7–16.
[32] D. Panov and J. Ross : Slope Stability and Exceptional Divisors of High Genus,
Math. Ann. 343 (2009), no. 1, 79–101. arXiv:0710.4078v1 [math.AG].
92 A. Futaki and Y. Sano
[33] S.T. Paul and G. Tian : CM Stability and the Generalized Futaki Invariant I.,
math.AG/0605278, 2006.
[34] S.T. Paul and G. Tian : CM Stability and the Generalized Futaki Invariant II. (To
appear in Asterisque), math.AG/0606505, 2006.
[35] D.H. Phong and J. Sturm : On stability and the convergence of the Kähler-Ricci
flow, J. Differential Geom. 72 (2006), no. 1, 149–168.
[36] D.H. Phong and J. Sturm : Lectures on stability and constant scalar curvature,
Handbook of geometric analysis, No. 3, 357–436, Adv. Lect. Math. (ALM), 14,
Int. Press, Somerville, MA, 2010. arXiv:0801.4179.
[37] D.H. Phong, N. Sesum and J. Sturm : Multiplier ideal sheaves and the Kähler-Ricci
flow, Comm. Anal. Geom. 15, no. 3 (2007), 613–632.
[38] J. Ross and R.P. Thomas : An obstruction to the existence of constant scalar
curvature Kähler metrics, J. Differential Geom. 72 (2006), no. 3, 429–466.
[39] Y.A. Rubinstein : The Ricci iteration and its applications, C. R. Acad. Sci. Paris,
Ser. I, 345 (2007), 445-448, arXiv:0706.2777.
[40] Y.A. Rubinstein : On the construction of Nadel multiplier ideal sheaves and the
limiting behavior of the Ricci flow, Trans. Amer. Math. Soc. 361 (2009), no. 11,
5839–5850. arXive:math/0708.1590.
[41] Y.A. Rubinstein : Some discretizations of geometric evolution equations and the
Ricci iteration on the space of Kähler metrics, Adv. Math. 218 (2008), no. 5,
1526–1565. arXive:math/0709.0990.
[42] Y. Sano : Multiplier ideal sheaves and the Kähler-Ricci flow on toric Fano mani-
folds with large symmetry, preprint, arXiv:0811.1455.
[43] N. Sesum and G. Tian : Bounding scalar curvature and diameter along the Kähler-
Ricci flow (after Perelman), J. Inst. Math. Jussieu 7 (2008), no. 3, 575–587.
[44] Y-T. Siu : The existence of Kähler-Einstein metrics on manifolds with positive
anticanonical line bundle and a suitable finite symmetry group, Annals of Math.
127 (1988) 585–627.
[45] Y-T. Siu : Dynamical multiplier ideal sheaves and the construction of rational
curves in Fano manifolds, Complex analysis and digital geometry, 323–360, Acta
Univ. Upsaliensis Skr. Uppsala Univ. C Organ. Hist., 86, Uppsala Universitet,
Uppsala, 2009. arXiv:0902.2809.
[46] J. Song : The α-invariant on toric Fano manifolds, Amer. J. Math. 127, No.6 (2005)
1247–1259.
[47] J. Stoppa : K-stability of constant scalar curvature Kähler manifolds, Adv. Math.
221 (2009), no. 4, 1397–1408. arXiv:0803.4095.
[48] G. Székelyhidi : The Kähler-Ricci flow and K-polystability, Amer. J. Math. 132
(2010), no. 4, 1077–1090. arXiv:0803.1613.
[49] G. Tian : On Kähler-Einstein metrics on certain Kähler manifolds with C1 (M) > 0,
Invent. Math. 89, no.2 (1987) 225–246.
[50] G. Tian, Kähler-Einstein metrics with positive scalar curvature, Invent. Math. 130,
no.1 (1997) 1–37.
[51] G. Tian and S.T. Yau, Kähler-Einstein metrics on complex surfaces with C1 (M)
positive, Comm. Math. Phys. 112 (1987) 175–203.
[52] G. Tian and X. Zhu : A new holomorphic invariant and uniqueness of Kähler-Ricci
solitons, Comment. Math. Helv 77, No.2 (2002) 297–325.
Multiplier ideal sheaves 93
[53] G. Tian and X. Zhu : Convergence of Kähler-Ricci flow, J. Amer. Math. Soc., 20,
No.3 (2007), 675–699.
[54] V. Tosatti : Kähler-Ricci flow on stable Fano manifolds, J. Reine Angew. Math.
640 (2010), 67–84. arXiv:0810.1895.
[55] X.-W. Wang : Moment maps, Futaki invariant and stability of projective manifolds,
Comm. Anal. Geom. 12 (2004), no. 5, 1009–1037.
[56] X.J. Wang and X. Zhu : Kähler-Ricci solitons on toric manifolds with positive first
Chern class, Adv. Math. 188, No.1 (2004) 87–103.
[57] B. Weinkove, A complex Frobenius theorem, multiplier ideal sheaves and
Hermitian-Einstein metrics on stable bundles, Trans. Amer. Math. Soc. 359, no.4
(2007) 1577–1592.
[58] S.-T.Yau : On the Ricci curvature of a compact Kähler manifold and the complex
Monge-Ampère equation I, Comm. Pure Appl. Math. 31(1978), 339–441.
[59] S.-T. Yau : Open problems in Geometry, Proc. Symp. Pure Math. 54 (1993) 1–28.
[60] R. Ye : The logarithmic Sobolev inequality along the Ricci flow, arXiv:0707.2424
(2007).
[61] O.S. Zhang : A uniform Sobolev inequality under Ricci flow, arXiv:0706.1594
(2007).
[62] X. Zhu, Kähler-Ricci soliton type equations on compact complex manifolds with
C1 (M) > 0, J. Geom. Anal., 10 (2000), 759–774.
Authors’ addresses:
Akito Futaki
Department of Mathematics,
Tokyo Institute of Technology, 2-12-1,
O-okayama, Meguro, Tokyo 152-8551,
Japan
[email protected]
Yuji Sano
Department of Mathematics,
Kyushu University,
6-10-1, Hakozaki, Higashiku,
Fukuoka-city, Fukuoka 812-8581
Japan
[email protected]
6
Multisymplectic formalism and the covariant
phase space
fr éd éric h élein
94
Multisymplectic formalism and the covariant phase space 95
open subset of X and consider the set C ∞ (U, Y) of smooth maps u from U to
Y. Let L : U × Y × End(X, Y) −→ R be a Lagrangian density and consider
the action functional on C ∞ (U, Y) defined by:
L[u] = L(x, u(x), dux )β,
U
μ
i ∂p
In fact this can be easily checked by choosing Xμ := ∂x∂ μ + ∂x μ ∂y i + ∂x μ ∂p μ .
∂u ∂ i ∂
i
More concisely we can introduce the n-multivector field X := X1 ∧ · · · ∧ Xn
(so that X(x) ∈ "n T(x,u(x),p∗ (x)) ∗ , ∀x ∈ U ). Then Equation (6.3) reads:
∀ξ ∈ X × Y × End(Y∗ , X∗ ),
μ
dpi ∧ dy i ∧ βμ (ξ, X) = dH (ξ )β(X). (6.4)
β|∗ = 0, (6.5)
ω := de ∧ β + dpiμ ∧ dy i ∧ βμ . (6.6)
2 W.l.g. we can assume that the constant h is zero, so that is included in H−1 (0).
98 F. Hélein
∂ a ui
vμi 1 ···μa j r u(x) = (x).
∂x μ1 · · · ∂x μa
It is convenient to introduce the multi-index notation M = μ1 · · · μa , where
a ∈ N and ∀b ∈ [[1, a]], 1 ≤ μb ≤ n and to set |M| = a. Then for |M| = r we
denote
∂L
πiM (x, v) := i
(x, v). (6.8)
∂vM
The analogue of the Legendre hypothesis consists here in supposing that the
map (vM i
)i,M;|M|=r −→ (piM )i,M;|M|=r defined by (6.8) is one to one. Next we
define the vector space M with coordinates:
where
∂ ∂
Dμ := μ
+ vMμ
i
i
.
∂x ∂vM
|M| < r. To any map u from U to Y we associate the map p∗ which is the
image of j r u by the maps πiM . Then u is a critical point of L iff (j r−1 u, p∗ ) is
Multisymplectic formalism and the covariant phase space 99
construction as for instance the example in [20, 35] obtained by starting from
the Palatini formulation of gravity.
One can then show that any n-dimensional submanifold on which β does
not vanish is a critical point of A iff it is a Hamiltonian n-curve, i.e. a solution
of (6.10) (see [18]). Actually if is the image of a given configuration by
some Legendre transform, then A[] coincides with the Lagrangian action of
the configuration we started with [19]. Note that in the case where ω is not
exact one could define a similar
action
on a homology class of n-dimensional
submanifolds by replacing by ω, where is a (n + 1)-chain connecting
with a particular n-dimensional submanifold which generates the homology
class.
dF + ξ ω = 0. (6.12)
102 F. Hélein
Then one can recover two important notions in the semi-classical theory of
fields. First one can define a bracket between observable (n − 1)-forms F and
G by the formula
of this equation in the quantization of fields is not clear for the moment. One of
the interests of the Hamilton–Jacobi equation is that it allows one in principle
to prove under some circumstances that some solutions of the Euler–Lagrange
system of equations are global minimizers, by following a classical strategy
designed by K. Weierstrass and D. Hilbert (see [47, 6, 36]). This strategy is the
exact analogue in the general theory of calculus of variations of the theory of
calibrations used in minimal surfaces.
Note that one could impose extra conditions such as requiring that λ =
ds 1 ∧ · · · ∧ ds n , where s 1 , . . . , s n are functions on Z plus the fact that the
graph of λ is foliated by solutions to the Hamilton equations (this provides then
a generalization of the picture built by Hamilton in order to reconcile the Fermat
principle with the Huygens principle): this was achieved by Carathéodory [2]
in his theory (see §6.1.8).
3 This was followed by a work by L. Koenigsberger [27] in 1901, quoted by T. De Donder in [8],
which unfortunately I have difficulties to understand.
Multisymplectic formalism and the covariant phase space 105
4 Including, in previous papers, the author of this note, who until recently was unaware of the
work of Volterra.
106 F. Hélein
6.1.9 An example
Let X be the n-dimensional Minkowski space-time with coordinates
x = (x 0 , x 1 , . . . , x n−1 )
ω := de ∧ β + dpμ ∧ dϕ ∧ βμ .
θλ := eβ + λpμ dϕ ∧ βμ − (1 − λ)ϕdpμ ∧ βμ ,
which satisfies
⎧
⎪
⎪ μν ∂ϕ
⎨ p μ (x) = η (x)
∂x ν
(6.17)
⎪
⎪ 1 ∂ϕ ∂ϕ 1
⎩ e(x) = − ημν (x) ν (x) − m2 ϕ(x)2 .
2 ∂x μ ∂x 2
We define E to be the set of Hamiltonian n-curves s.t. for all time t, x −→
ϕ(t, x) is rapidly decreasing at infinity.
We denote by Pn−1 H M the set of dynamical observable (n − 1)-forms
F and spH M := {ξ | Lξ ω = 0, dH(ξ ) = 0}. Note that (n − 1)-forms F in
H M are found by looking at vector fields ξ in spH M and by solving
Pn−1
Multisymplectic formalism and the covariant phase space 107
(We remark that the maps { Pλ(λ) , ·} play the role of the generators of a Cartan
subalgebra.) Let us denote by C the mass shell, i.e. the set of all k = (k 0 , k) ∈
R4 which are solutions of (6.20). This set actually splits into two connected
components according to the sign of k 0 : we let C + := {k ∈ C | k 0 > 0}. For any
k ∈ C + we define
i
αk := Fieik·x /√2π 3 = √ 3
eik·x (pμ − iϕk μ ) βμ
2π
−i −ik·x μ
αk∗ := F−ie−ik·x /√2π 3 = √ 3e (p + iϕk μ ) βμ .
2π
The vector fields associated to these observable forms are:
/ 0
eik·x ∂ μ ∂
∂
ξk := ξαk = √ 3 i −k + ημν p k − im ϕ
μ ν 2
,
2π ∂ϕ ∂pμ ∂e
/ 0
e−ik·x ∂ ∂ ∂
ξk∗ := ξαk∗ = √ 3 −i − k μ μ + ημν pμ k ν + im2 ϕ .
2π ∂ϕ ∂p ∂e
As the notations suggest these functionals are the classical analogues of respec-
tively the annihilation and the creation operators. The advantage however is
that our functionals ak and ak∗ are independent of the coordinate system. We
can choose to be the hyperplane x 0 = t = 0 and, for any function f , denote
by f |0 the restriction of f to . Then, for any ∈ E we have
1
i ∂ϕ ∂ϕ 20 (k),
ak () = √ 3
|0 (x) − ik 0 ϕ|0 (x) e−ik·x d x = i |0 (k) + k 0 ϕ|
2π R3 ∂t ∂t
Similarly we have:
−i ∂ϕ
ak∗ () = √ 3 |0 (x) + ik 0 ϕ|0 (x) eik·x d x
2π R ∂t
3
1
∂ϕ 20 (−k).
= −i |0 (−k) + k 0 ϕ|
∂t
Multisymplectic formalism and the covariant phase space 109
A way to regularize these operators and their brackets is, by using functions
f, g ∈ L2 (C + ), to define
af := dμ(k)f (k)ak , and ag∗ := dμ(k)g(k)ak∗ .
C+ C+
Then
{af , ag∗ } = i dμ(k)f (k)g(k).
C+
(M, ω, β) (see §6.1.4) and, as in §6.1.5, that ω is exact, i.e. ω = dθ , for some
n-form θ . We note E the set of Hamiltonian n-curves in (M, ω, β), i.e. the set
of oriented n-dimensional submanifolds ⊂ M which satisfy (6.10). Given
some ∈ E, the tangent space6 to E at represents the set of infinitesimal
deformations δ of which preserves the equation (6.10). Such a deformation
δ can be represented by a vector field ξ tangent to M defined along , i.e. a
section over of j∗ T M, which is the pull-back image of the tangent bundle
T M by the embedding map j : −→ M. Given δ, the vector field ξ is of
course not unique, since for any tangent vector field ζ on (i.e. a section of
the subbundle T ⊂ j∗ T M), ξ + ζ represents also δ. If so we write:
δ = ξ= ξ + ζ.
The dependence of on
This is the first natural question. For that purpose let us consider a smooth
1-parameter family of slices (t )t and compute the derivative:
d t d
(δ) = ξ θ = L ∂t∂ (ξ θ )
dt dt t ∩ t ∩
∂ ∂
= d(ξ θ ) + d ξ ∧ θ .
t ∩ ∂t ∂t
But d(ξ θ ) = Lξ θ − ξ dθ and thus
∂ ∂ ∂
d(ξ θ) = Lξ θ − ξ ω.
∂t ∂t ∂t
6 Note that since E may not be a manifold in general, the usual definition of a tangent space
should be replaced by a suitable notion, see [25, 18].
Multisymplectic formalism and the covariant phase space 113
∂ ∂ ∂
d(ξ θ ) = Lξ θ −ξ ∧ ω.
∂t ∂t ∂t
Hence
d t ∂ ∂
(δ) = Lξ θ − ξ∧ ω
dt t ∩ ∂t t ∩ ∂t
∂
+ d ξ∧ θ . (6.22)
t ∩ ∂t
δ
(δ1 , δ2 ) = δ1 ·
(δ2 ) − δ2 · (δ1 )
− ([δ1 , δ2 ])
= δ1 · ξ2 θ − δ2 · ξ1 θ .
∩ ∩
Thus
δ
(δ1 , δ2 ) = Lξ1 (ξ2 θ) − Lξ2 (ξ1 θ) .
∩ ∩
We use then the following identity (see [18]): for any pair of vector fields X1
and X2 and for any p-form β,
Γ+
∑1
σ2
Γ0 Γ0’
∑2
σ1
Γ–
1 2
Figure 6.1 A geometric comparison of (δ) with (δ)
by 0 the part of which is between 1 and 2 (see again the picture). Lastly
we consider the (not necessarily Hamiltonian) n-curve ε , which is the union
of − , σ1 , 0 , σ2 and + . Of course ε is not smooth, but it can be approached
by a sequence of smooth n-curves, so that the following makes sense. We also
endow ε with the orientation which agrees with that of on − ∪ + and
with that on 0 .
Let us use the fact that is a Hamiltonian n-curve, hence a critical point of
(6.11). This implies that
θ= θ + o(ε). (6.27)
ε
We now recognize that, on the one hand, σ1 θ = ε 1 (δ) + o(ε), σ2 θ =
−ε 2 (δ) + o(ε) (the sign
being due to the orientation of σ2 ). On the other
hand 0 θ = S12 (0 ) and θ = S12 (0 ) = S12 (0 + εδ0 ) + o(ε). Hence
0
(6.28) gives us
ε 1
(δ) + ε δS12 (δ) − ε 2
(δ) = o(ε).
Spacetime translations
We now look at the canonical vector fields on E associated with spacetime
translations Pζ , where ζ is constant vector field on X . We recall that ξP (λ) = ζ .
ζ
We must understand the induced vector field ζ on F. Let U (s, ·) be the flow
mapping of the vector field ζ : U (s, x, ϕ, e, p) = (s, x + ζ, ϕ, e, p). Then the
image of
by U (x, ·) is
where the value of es (x) and ps (x) is completely determined by the constraint
that s ⊂ E and by the knowledge of ϕ(x − sζ ). This can be proved by a simple
change of variable. Similarly we determine the action of ζ on the coordinates
(uk , u∗k )k∈C + by computing its action on ϕ:
!
−ik·(x−sζ )
d 1
ζ ϕ (x) = √ 3 + dμ(k) uk e + u∗k eik·(x−sζ )
ds 2π C
s=0
1
= √ 3 dμ(k) ik · ζ uk e−ik·(x−sζ ) − ik · ζ u∗k eik·(x−sζ )
+
2π C
/ 0
∂ ∗ ∂
=i k · ζ uk − uk ∗ ϕ (x).
C+ ∂uk ∂uk
Hence
∂ ∂
ζ = i k · ζ uk − u∗k ∗ .
C+ ∂uk ∂uk
Geometric prequantization
We recall very briefly the prequantization scheme due to B. Kostant and J.-M.
Souriau (generalizing previous constructions by B.O. Koopman, L. Van Hove
and I. Segal, see [26, 38]). We let (M, ω) be a simply connected symplectic
120 F. Hélein
manifold and we assume for simplicity that there exists a 1-form θ with ω = dθ .
We consider the trivial bundle L := M × C and denote by (M, L) the set of
square integrable sections of L. Using θ we can define a Hermitian connection
∇ acting on (M, L) by
i
∀ξ ∈ (M, T M), ∀ψ ∈ (M, L), ∇ξ ψ = ξ · ψ − θ (ξ )ψ.
3 acting on
Then to each function F ∈ C ∞ (M, R) we associate the operator F
(M, L)
3ψ = F ψ + ∇ξF ψ = (F − θ (ξF )) ψ + ξF · ψ,
F
i i
where dF + ξF ω = 0. This construction is called the prequantization of
(M, ω). For instance if (M, ω) = (R2n , dpi ∧ dq i ), then ω = dθ , with θ =
pi dq i and q3i = q i + i ∂p∂ i and p
3i = −i ∂q∂ i . Of course one needs further
restrictions in order to recover an irreducible representation of the Heisenberg
algebra (and hence the standard quantization): this will be the purpose of
introducing a polarization and a tensorization of the line bundle L with the
bundle of half volume forms transversal to the leaves of the polarization (see
[26, 38]).
We will propose an extension of this procedure to our setting, concerned
with the quantization of fields. We consider the trivial bundle L := E C × C
over E C , where E C is the complexification of the set of solutions to the Klein–
Gordon equation (6.16) as before. On the set (E C , L) of smooth sections of
L (we are here relatively vague about the meaning of "smooth") we define a
notion of covariant derivative along any vector field of the type F , where
F ∈ (E C , L) by
C i
∀ψ ∈ (E , L), ∇F ψ := F · ψ − ξF θ ψ,
then
uk + u∗k
1
√ 3 e−ik·x ϕ(0, x)d x = ,
2π Rn 2k 0
and
uk − u∗k
1
√ 3 e−ik·x p0 (0, x)d x = ,
2π Rn 2i
Similarly
dk
αg∗ = g(k)u∗k .
∩ C+ 2k 0
αg∗
and similarly ξg∗ θ= 2
. Hence
d k f (k) d k g(k) ∗
ξf θ= uk and ξg∗ θ= u .
∩ C+ 2k 0 2 ∩ C+ 2k 0 2 k
Using the previous results we can now express, for ψ ∈ (E C , L),
i
∇f ψ := f · ψ − ξf θ ψ
∂ψ i d k f (k)
= i f (k) ∗ − uk ψ
C+ ∂uk C + 2k 0 2
122 F. Hélein
and
i
∇∗g ψ := ∗g · ψ − ξ∗ θ ψ
g
∂ψ i d k g(k) ∗
= −i g(k) − uk ψ.
C + ∂uk C + 2k 0 2
and
∂ψ d k g(k) ∗
ag∗ ψ = −
3 g(k) + u ψ.
∂uk C+ 2k 0 2 k
dk
[3 ag∗ ] =
af ,3 f (k)g(k).
C+ 2k 0
∂ ∂
P1
(λ) ψ =
ζ ζ · ψ = k · ζ uk − u∗k ∗ ψ.
i C+ ∂uk ∂uk
Note that if we need to compute Pζ(λ) , it is more suitable to set λ = 1, since
it gives then the standard expression for the stress-energy tensor. For instance,
if ζ = ∂x∂ 0
3
!
(p 0 )2 (pi )2 ϕ2
− P0(1) = dx + + m2
∩ Rn−1 2 i=1
2 2
gives the total energy in the frame associated with the coordinates x μ .
Multisymplectic formalism and the covariant phase space 123
Introducing a polarization
We choose to impose the extra condition ∇∗g ψ = 0, ∀g (covariantly antiholo-
morphic sections), which gives us:
1 dk
ψ(uk , u∗k ) = h(u∗k ) exp − uk u∗k = h(u∗k )|0.
2 C + 2k 0
The advantage of this choice is that all observable functionals (creation and
annihilation, energy and momentum) are at most linear in the variables (uk , u∗k ),
so we do not need to use the Blattner–Kostant–Sternberg correction for these
3|0 = 0, so that the energy of the vacuum van-
operators [26, 38]. As a result P
ishes without requiring normal ordering. However we did not take into account
the metaplectic correction, which requires a slight change of the connection:
were we to do so we would find that the vacuum has infinite energy (as in
the standard quantization scheme), which can be removed by a normal order-
ing procedure. The mysterious thing here (as was already observed) is that by
ignoring the metaplectic correction (which however is fundamental for many
reasons) we do not need the normal ordering correction.
References
[1] J. C. Baez, C. L. Rogers, Categorified Symplectic Geometry and the String Lie
2-algebra, Homology, Homotopy Appl. 12 (2010), no. 1, 221–236. arXiv:0901.
4721.
[2] C. Carathéodory, Variationsrechnung und partielle Differentialgleichungen erster
Ordnung, Teubner, Leipzig (reprinted by Chelsea, New York, 1982); Acta litt. ac
scient. univers. Hungaricae, Szeged, Sect. Math., 4 (1929), p. 193.
[3] E. Cartan, Leçons sur les invariants intégraux, Hermann, 1922.
[4] A. Clebsch, Ueber die zweite Variation vielfache Integralen, J. reine angew. Math.
56 (1859), 122–148.
[5] C. Crnkovic, E. Witten, Covariant description of canonical formalism in geo-
metrical theories, in Three hundred years of gravitation (S. W. Hawking and
W. Israel, eds.), Cambridge University Press, Cambridge, 1987, 676–684; E.
Witten, Interacting field theory of open supertrings, Nucl. Phys. B 276 (1986),
291–324.
[6] P. Dedecker, Calcul des variations, formes différentielles et champs géodésiques,
in Géométrie différentielle, Colloq. Intern. du CNRS LII, Strasbourg 1953, Publ.
du CNRS, Paris, 1953, p. 17–34; On the generalization of symplectic geometry
to multiple integrals in the calculus of variations, in Differential Geometrical
Methods in Mathematical Physics, eds. K. Bleuler and A. Reetz, Lect. Notes
Maths. vol. 570, Springer-Verlag, Berlin, 1977, p. 395–456.
[7] T. De Donder, Introduction à la théorie des invariants intégraux, Bull. Acad. Roy.
Belgique (1913), 1043–1073.
124 F. Hélein
Mechanik für mehrere unavhängige Variable, J. Reine Angw. Math., Bd. 124
(1902), 202–277.
[28] Y. Kosmann-Schwarzbach, Les théorèmes de Noether – Invariance et lois de
conservation au XXe siècle, Les éditions de l’Ecole Polytechnique, 2004.
[29] T. Lepage, Sur les champs géodésiques du calcul des variations, Bull. Acad. Roy.
Belg., Cl. Sci. 27 (1936), 716–729, 1036–1046.
[30] E. Noether, Invariante Variationsprobleme, Nachrichten von der Königlichen
Gesellschaft des Wissenschaften su Göttingen, Mathematisch-physikalische
Kalsse, 1918, p. 235–257.
[31] R.E. Peierls, The commutation laws of relativistic field theory, Proc. Roy. Soc.
London, Ser. A, Vol. 214, No. 1117 (1952), 143–157.
[32] G. Prange, Die Hamilton–Jacobische Theorie für Doppelintegrale, Diss.
Göttingen, 1915.
[33] H. Poincaré, Les méthodes nouvelles de la mécanique céleste, t. III, Paris, Gauthier–
Villars, 1899.
[34] E.G. Reyes, On covariant phase space and the variational bicomplex, Int. J. Theor.
Phys., Vol. 43, No. 5 (2004), 1267–1286.
[35] C. Rovelli, A note on the foundation of relativistic mechanics – II: Covari-
ant Hamiltonian general relativity, in Topics in Mathematical Physics, Gen-
eral Relativity and Cosmology, H Garcia-Compean, B Mielnik, M Montesinos,
M Przanowski editors, pg 397, (World Scientific, Singapore 2006). arXiv:gr-
qc/0202079
[36] H. Rund, The Hamilton–Jacobi theory in the calculus of variations, its role in
mathematics and physics, Krieger Pub. 1973.
[37] I. Segal, Quantization of nonlinear systems, J. Math. Phys. vol. 1, N. 6 (1960),
468–488.
[38] J. Śnyatycki, Geometric quantization and quantum Mechanics, Appl. Math. Sci.
30, Springer-Verlag 1980.
[39] J.-M. Souriau, Structure des systèmes dynamiques, Dunod, Paris, 1970.
[40] F. Takens, A global formulation of the inverse problem of the calculus of variations,
J. Diff. Geom. 14 (1979), 543–562.
[41] A.M. Vinogradov, The C-spectral sequence, Lagrangian formalism, and conser-
vations laws, I and II, J. Math. Anal. Appl. 100 (1984), 1–40 and 41–129.
[42] L. Vitagliano, Secondary calculus and the covariant phase space, J. Geom. Phys.
59 (2009), no. 4, 426–447.
[43] L. Vitagliano, The Lagrangian–Hamiltonian Formalism for Higher Order Field
Theories, J. Geom. Phys. 60 (2010), no. 6-8, 857–873. arXiv:0905.4580.
[44] L. Vitagliano, Partial Differential Hamiltonian Systems, preprint arXiv:0903.4528
[45] V. Volterra, Sulle equazioni differenziali che provengono da questiono di calcolo
delle variazioni, Rend. Cont. Acad. Lincei, ser. IV, vol. VI, 1890, 42–54.
[46] V. Volterra, Sopra una estensione della teoria Jacobi–Hamilton del calcolo delle
varizioni, Rend. Cont. Acad. Lincei, ser. IV, vol. VI, 1890, 127–138.
[47] H. Weyl, Geodesic fields in the calculus of variation for multiple integrals, Ann.
Math. (3) 36 (1935), 607–629.
[48] G. Zuckerman, Action functional and global geometry, in Mathematical aspects
of string theory, S.T. Yau, eds., Advanced Series in Mathematical Physics, vol 1,
World Scientific, 1987, 259–284.
126 F. Hélein
Author’s address:
7.1 Introduction
The search for manifolds of nonnegative curvature1 is one of the classical
problems in Riemannian geometry. While general obstructions are scarce, there
are relatively few general classes of examples and construction methods. Hence,
it is unclear how large one should expect the class of closed manifolds admitting
a nonnegatively curved metric to be. For a survey of known examples, see
e.g. [12].
Apart from taking products, there are only two general methods to construct
new nonnegatively curved metrics out of given spaces. One is the use of Rie-
mannian submersions which do not decrease curvature by O’Neill’s formula.
The other is the glueing of two manifolds (which we call halves) along their
common boundary. Typically, the boundary of each half is assumed to be totally
geodesic or, slightly more restrictively, a collar metric. This in turn implies by
the Soul theorem ([2]) that each half is the total space of a disk bundle over a
totally geodesic closed submanifold. In addition, the glueing map of the two
boundaries must be an isometry.
While many examples can be constructed by such a glueing, its application
is still limited. On the one hand, there is not too much known on the question
of which disk bundles over a nonnegatively curved compact manifold admit
collar metrics of nonnegative curvature, and on the other hand, even if such
metrics exist, the metric on the boundary is not arbitrary. Thus, glueing together
two such disk bundles to a nonnegatively curved closed manifold is possible in
special situations only.
For instance, if the disk bundle is homogeneous, then there always exist
invariant nonnegatively curved collar metrics. However, the metric on the
1 Throughout this article, the term “curvature” refers to the sectional curvature.
127
128 L. J. Schwachhöfer
If H is the trivial group so that G/H = G, then (7.2) in the above theorem
is equivalent to a simpler criterion. Namely, we have the following
Corollary 7.4.2 ([1]) Let M1n and M2n be CROSSes of equal dimension. Then
both M1 #M2 and M1 #M 2 admit nonnegatively curved metrics.
∂(M1 × M2 ).
Example 7.5.3
(i) Z2 -quotients. Let K K ⊂ G be such that K /K ∼= Z2 , and let L →
O(V ) be a representation which is transitive on the unit sphere. More-
over, define the representation K → K /K ∼
= {±I d} ⊂ O(V ). Then the
essentially trivial bundle
(G × L) ×K ×L V ∼
= G ×K V −→ G/K
(G × Z2 ) ×K ×Z2 V ∼
= G ×K V → G/K ,
(G × L) ×K ×L V ∼
= G ×K V −→ G/K
(G × L) ×K ×L V ∼
= G ×K V −→ G/K
Theorem 7.5.5 ([9]) Let H ⊂ K ⊂ G be compact Lie groups with Lie alge-
bras h ⊂ k ⊂ g, and consider the decompositions from (7.1). If there exists
C > 0 such that for all X = Xm + Xs , Y = Y m + Y s ∈ m ⊕ s = p we have
That is, for all Lie groups H ⊂ K ⊂ G satisfying this criterion, the metric
gg from (7.5) has nonnegative curvature on all planes contained in p for small
ε > 0, hence if K/H = S V , then the corresponding homogeneous vector bun-
dle G ×K V → G/K has a nonnegatively curved normal homogeneous collar
metric.
As it turns out, condition (7.6) is almost necessary and sufficient for a
homogeneous disk bundle to admit such a metric. Namely, we have
To describe our results for rank three and four bundles, we require some
notation for subgroups of the exceptional Lie group G2 . Let SO(4) ⊂ G2
and SU (3) ⊂ G2 be the isotropy groups of the symmetric space G2 /SO(4)
and the sphere S 6 = G2 /SU (3), respectively. After conjugating these groups
appropriately, their intersection can be made isomorphic to U (2), and we
let SU (2)1 ⊂ SO(4) ∩ SU (3) ⊂ G2 be the simple part of this intersection.
Note that SU (2)1 ⊂ SO(4) is normal, and we denote its centralizer in G2
by SU (2)3 ⊂ SO(4) ⊂ G2 . (The subscripts of the SU (2)-subgroups of SO(4)
denote their maximal weight for the isotropy representation of G2 /SO(4).)
Using this notation, we can make the following statement about the rank-three
case.
(i) M1 = G2 ×SO(4) su(2)3 , where SO(4) acts on su(2)3 so(4) by the adjoint
representation.
(ii) M2 = (Sp(p + 1) · G ) ×Sp(1)·H sp(1) with H ⊂ Sp(p) · G , where
Sp(1) · H acts on sp(1) sp(1) ⊕ h by the adjoint representation.
Finally, in the rank-four case, we have the following examples, which are
all related to those in Theorem 7.5.10.
(i) G2 ×SO(4) (H/± 1), where SU (2)1 ⊂ SO(4) acts trivially and SU (2)3 ⊂
SO(4) by left multiplication on H/± 1. Note that this is an orbifold bundle
only.
(ii) (G2 × G ) ×SO(4)×SU (2) (H/± 1), where SU (2)1 ⊂ SO(4) acts trivially
and SU (2)3 ⊂ SO(4) by left multiplication, whereas SU (2) ⊂ G acts
by right multiplication on H/± 1 with G arbitrary. Note that these are
orbifold bundles only.
(iii) Sp(p + 1) ×Sp(1)×Sp(p) H where Sp(p) acts trivially and Sp(1) by left
multiplication on H.
Nonnegative curvature on disk bundles 139
M = D− ∪G/H D+ ,
Theorem 7.6.2 ([5]) Let M be a cohomogeneity one manifold with two non-
principal orbits of codimension at most 2. Then M admits a G-invariant metric
of nonnegative curvature.
This result yields some spectacular examples of manifolds which admit non-
negatively curved metrics, such as all four (unoriented) diffeomorphism types
of RP5 and ten out of the fourteen (unoriented) exotic spheres of dimension 7.
Unfortunately, the remaining homogeneous disk bundles admitting nonneg-
atively curved normal homogeneous collar metrics do not yield new examples
of nonnegatively curved cohomogeneity one manifolds, since on glueing two
of them, we obtain either the restriction of a homogeneous action on M, or
there is a Riemannian submersion M̂ → M, where M̂ is known to have a
nonnegatively curved metric.
For instance, by glueing together two disk bundles of the second type of
Theorem 7.5.8 with p = 0, after applying outer automorphisms of Spin(8), we
conclude that the primitive cohomogeneity one manifold given by the group
diagram G2 ⊂ {Spin+ (7), Spin− (7)} ⊂ Spin(8) admits a metric of nonneg-
ative curvature with a totally geodesic normal homogeneous principal orbit.
However, this manifold is diffeomorphic to the sphere S 15 ([4]).
References
[1] J. Cheeger, Some examples of manifolds of nonnegative curvature, J. Diff. Geom.
8 (1973) 623–628
[2] J. Cheeger, D. Gromoll, On the structure of complete manifolds of nonnegative
curvature, Ann. of Math. 96 (1972) 413–443.
[3] K. Grove, L. Verdiani, B. Wilking, W. Ziller, Nonnegative curvature obstructions
in cohomogeneity one and the Kervaire spheres, Ann. Sc. Norm. Super. Pisa, Cl.
Sci. (5) 5, No. 2 (2006) 159–170
[4] K. Grove, B. Wilking, W. Ziller, Positively curved cohomogeneity one manifolds
and 3-Sasakian geometry, J. Diff. Geom. 78 (2008) 33–111.
Nonnegative curvature on disk bundles 141
[5] K. Grove, W. Ziller, Curvature and symmetry of Milnor spheres, Ann. of Math.
(2) 152 (2000) 331–367.
[6] L. Guijarro, Improving the metric in an open manifold with nonnegative curvature,
Proc. Amer. Math. Soc., 126 (1998) 1541–1545
[7] G. Perelman, Proof of the soul conjecture of Cheeger and Gromoll, J. Diff. Geom.
40 (1994), 209–212.
[8] L. Schwachhöfer, A remark on left invariant metrics on compact Lie groups,
Arch.Math. 90 (2008) 158–162
[9] L. Schwachhöfer, K. Tapp, Homogeneous Metrics with nonnegative curvature,
Jour. Geom. Anal. 19 (2009) 929–943
[10] L. Schwachhöfer, K. Tapp, Cohomogeneity one disk bundles with normal homo-
geneous collars, Proc. London Math. Soc. 99, (2009) 609–632
[11] L. Schwachhöfer and W. Tuschmann, Almost nonnegative curvature and coho-
mogeneity one, Preprint no. 62/2001, Max-Planck-Institut für Mathematik in den
Naturwissenschaften Leipzig, http://www.mis.mpg.de/cgi-bin/preprints.pl
[12] W. Ziller, Examples of Riemannian manifolds with nonnegative sectional curva-
ture, Metric and Comparison Geometry, Surv. Diff. Geom. 11, ed. K. Grove and
J. Cheeger, Intern. Press, 2007
Author’s address:
Abstract
We study the Morse theory of the Yang-Mills-Higgs functional on the space
of pairs (A, #), where A is a unitary connection on a rank 2 hermitian vector
bundle over a compact Riemann surface, and # is a holomorphic section
of (E, dA ). We prove that a certain explicitly defined substratification of the
Morse stratification is perfect in the sense of G-equivariant cohomology, where
G denotes the unitary gauge group. As a consequence, Kirwan surjectivity holds
for pairs. It also follows that the twist embedding into higher degree induces
a surjection on equivariant cohomology. This may be interpreted as a rank 2
version of the analogous statement for symmetric products of Riemann surfaces.
Finally, we compute the G-equivariant Poincaré polynomial of the space of τ -
semistable pairs. In particular, we recover an earlier result of Thaddeus. The
analysis provides an interpretation of the Thaddeus flips in terms of a variation
of Morse functions.
8.1 Introduction
In this paper we revisit the notion of a stable pair on a Riemann surface. We
introduce new techniques for the computation of the equivariant cohomology
of moduli spaces. The main ingredient is a version of Morse theory in the spirit
of Atiyah and Bott [1] adapted to the singular infinite dimensional space of
holomorphic pairs.
Recall first the basic idea. Let E be a hermitian vector bundle over a closed
Riemann surface M of genus g ≥ 2. The space A(E) of unitary connections on
E is an infinite dimensional affine space with an action of the group G of unitary
gauge transformations. Via the Chern connection there is an isomorphism
A → dA between A(E) and the space of (integrable) Dolbeault operators (i.e.
142
Morse theory and stable pairs 143
Theorem 8.1.1 (Equivariantly perfect stratification) For every τ , d/2 < τ <
d, there is a G-invariant stratification of B(E) defined via the Yang-Mills-Higgs
flow that is perfect in G-equivariant cohomology.
The fact that perfection fails for the Morse stratification but holds for a
substratification seems to be a new phenomenon. In any case, as with vector
bundles, Theorem 8.1.1 allows us to compute the G-equivariant cohomology
144 R. A. Wentworth and G. Wilkin
prove the perfection of the stratification at this step, and the crucial Morse-
Bott lemma (Theorem 8.3.18), for d > 4g − 4. This we do in Section 8.3.5.
By contrast, for the other critical strata there is no such requirement on the
degree. Using this fact, we then give an inductive argument by twisting E
by a positive line bundle and embedding B(E) into the space of pairs for
higher degree, thus indirectly concluding the splitting of the associated long
exact sequence even at minimal Yang-Mills connections in low degree (see
Section 8.3.7).
This line of reasoning leads to another interesting consequence. For τ close
to d/2, there is a surjective holomorphic map from Mτ,d to the moduli space
of semistable rank 2 bundles of degree d. This is the rank 2 version of the
Abel-Jacobi map [4]. In this sense, Mτ,d is a generalization of the d-th sym-
metric product S d M of M. Choosing an effective divisor on M of degree k,
there is a natural inclusion S d M → S d+k M, and it was shown by MacDonald
in (14.3) of [16] that this inclusion induces a surjection on rational cohomol-
ogy. A similar construction works for rank 2 pairs, except now d → d + 2k,
while there is also a shift in the parameter τ → τ + k. We will prove the
following.
Remark 8.1.4 It is also possible to construct a moduli space of pairs for which
the isomorphism class of det E is fixed, indeed this is the space studied by
Thaddeus in [20]. The explicit calculations in this paper are all done for the
non-fixed determinant case, however it is worth pointing out here that the
idea is essentially the same for the fixed determinant case, and that the only
major difference between the two cases is in the topology of the critical sets.
In particular, the indexing set τ,d for the stratification is the same in both
cases, and Theorems 8.1.1 and 8.1.2 hold for the fixed determinant spaces as
well.
Definition 8.2.1 For a stable holomorphic bundle E, set μ+ (E) = d/2. For E
unstable, let
Let
Bss
τ
= (A, #) ∈ B : ((E, dA ), #) is τ -semistable
Morse theory and stable pairs 147
δ ∈ +
τ,d ⇐⇒ δ = τ − μ− (E, #), for some pair (E, #)
δ ∈ −
τ,d ⇐⇒ δ = μ+ (E) − τ, for some pair (E, #)
(iii), 0 ≤ deg # < d − τ implies τ < μ− (E, #). The last statement is clear,
since τ > d/2 implies τ − μ− (E, #) = μ+ (E) − d + τ > μ+ (E) − τ .
Corollary 8.2.6 −
τ,d ⊂ (0, d − τ ].
Remark 8.2.7 If δ ∈ + τ,d and δ < τ − d/2, then δ ≤ τ − d/2 − 1/2. Indeed,
if δ + d − τ = k ∈ Z, the condition forces k < d/2; hence, k ≤ d/2 − 1/2.
0 −→ L1 −→ E −→ L2 −→ 0 (8.2)
are extensions
0 −→ L2 −→ E −→ L1 −→ 0
For simplicity of notation, when τ is fixed we will mostly omit the superscript:
Bδ = Bδτ .
Remark 8.2.8 It is simple to verify that the stratification obtained above coin-
cides with the possible Harder-Narasimhan filtrations of pairs (E, #) consid-
ered as coherent systems (see [15, 18, 12]).
Morse theory and stable pairs 149
For δ ∈ +
τ,d ∩ Iτ,d , let
, ,
Xδ = Bδ ∪ Aj (δ )
δ ≤δ , δ ∈τ,d δ <δ , δ ∈+
τ,d ∩Iτ,d
For δ ∈ +
τ,d ∩ Iτ,d , let
, ,
Xδ = Bδ ∪ Aj (δ )
δ <δ , δ ∈ τ,d δ <δ , δ ∈+
τ,d ∩Iτ,d
Note that Xδ1 ⊂ Xδ Xδ ⊂ Xδ 2 , where δ1 is the predecessor and δ2 is the
successor of δ in τ,d . If δ ∈ +
τ,d ∩ Iτ,d , then Xδ = Xδ1 and Xδ = Xδ2 . In the
special case δ = τ − d/2, we have
Proposition 8.2.10 The sets {Xδ , Xδ }δ∈τ,d , are locally closed in B, G-
invariant, and satisfy
, ,
B= Xδ = Xδ
δ∈τ,d δ∈τ,d
, ,
Bδ ⊂ Bδ = Bδ ∪ Bδ
δ≤δ , δ ∈ τ,d δ<δ , δ ∈τ,d
, ,
Bδ ⊂ Bδ = Bδ ∪ Bδ
δ≤δ , δ ∈τ,d δ<δ , δ ∈τ,d
C(A,#)
0 D1
/ C1 D2
/ C2
(A,#) (A,#)
D1
/ 0,1 (End E) ⊕ 0 (E) D2
/ 0,1 (E) (8.8)
0 (End E)
D1 (u) = (−dA u, u#) , D2 (a, ϕ) = dA ϕ + a#
Here, D1 is the linearization of the action of the complex gauge group G C on
B, and D2 is the linearization of the condition dA # = 0. Note that D2 D1 = 0
if (A, #) ∈ B. The hermitian metric gives adjoint operators
D1∗ (a, ϕ) = −(dA )∗ a + ϕ#∗ , D2∗ (β) = (β#∗ , (dA )∗ β) (8.9)
The spaces of harmonic forms are by definition
H0 (C(A,#) ) = ker D1
H1 (C(A,#) ) = ker D1∗ ∩ ker D2
H2 (C(A,#) ) = ker D2∗
Vectors in 0,1 (End E) ⊕ 0 (E) that are orthogonal to the G C -orbit through
(A, #) are in ker D1∗ , and a slice for the action of G C on B is therefore given by
S(A,#) = ker D1∗ ∩ (a, ϕ) ∈ 0,1 (End E) ⊕ 0 (E) : D2 (a, ϕ) + aϕ = 0
(8.10)
1 More generally, the scale invariant parameter is τ vol(M)/2π .
152 R. A. Wentworth and G. Wilkin
: (ker D1 )⊥ × S(A,#) → B
(8.11)
(u, a, ϕ) → eu · (A + a, # + ϕ)
Proof By Lemma 8.2.15 and Proposition 8.2.14, the Kuranishi map gives a
homeomorphism of the slice with H1 (C(A,#) ). Hence, it suffices to define the
retraction there. For this we take
rt (a11 , a12 , a21 , a22 ; ϕ1 , ϕ2 ) = (ta11 , ta12 , a21 , ta22 ; tϕ1 , ϕ2 ) , t ∈ [0, 1]
For a sufficiently small neighborhood of the origin in the slice, this preserves
the set S(A,0) . It is also clearly equivariant.
154 R. A. Wentworth and G. Wilkin
fτ (A, #) = μ + iτ · id 2
(8.1)
μ(A, #) + iτ · id = 0 (8.2)
μ + iτ · id = ∗FA − i##∗ + iτ · id = 0
i (8.3)
=⇒ Tr(∗FA − i##∗ ) = 2τ ⇐⇒ deg E + # 2
= 2τ
2π M
Therefore 2τ ≥ d (with strict inequality if we want to ensure that # = 0).
Theorem 2.1.6 of [3] shows that a solution to the τ -vortex equations which is
not τ -stable must split. Moreover, since rankE = 2 the solutions can only split
if τ is an integer. In particular, for a generic choice of τ solutions to (8.2) must
be τ -stable. In general, critical sets of fτ can be characterized in terms of a
decomposition of the holomorphic structure of E. The critical point equations
for the functional fτ are
Let S d M denote the d-th symmetric product of the Riemann surface M, and
Jd (M) the Jacobian variety of degree d line bundles on M. For future reference
we record the following
HG∗ (ηδ ) =
⎧
⎪
⎪ HG∗ (Ass ) Type I, δ = τ − d/2
⎪
⎪
⎨H ∗ (J (M) × J ∗
j (δ) d−j (δ) (M)) ⊗ H (BU (1) × BU (1)) Type I, δ = τ − d/2
⎪H ∗ (S j (δ) M × Jd−j (δ) (M)) ⊗ H ∗ (BU (1))
⎪ Type II+
⎪
⎪
⎩ ∗ d−j (δ)
H (S M × Jj (δ) (M)) ⊗ H ∗ (BU (1)) Type II−
A standard calculation (cf. [3, Section 4]) shows that fτ can be re-written as
2 2
fτ = |FA |2 + dA # + ##∗ − 2τ |#|2 + |τ |2 dvol + 4τ deg E
X
(8.7)
156 R. A. Wentworth and G. Wilkin
This is very similar to the functional YMH studied in [10], and the proof
for existence of the flow for all positive time follows the same structure (which
is in turn modeled on Donaldson’s proof for the Yang-Mills functional in [9]),
therefore we omit the details. An important part of the proof worth mentioning
here is that the flow is generated by the action of G C , i.e. for all t ∈ [0, ∞)
there exists g(t) ∈ G C such that the solution (A(t), #(t)) to the flow equations
(8.6) with initial condition (A, #) is given by (A(t), #(t)) = g(t) · (A, #).
To show that the gradient flow converges, one can use the results of Theorem
B of [11] (where again, the functional is not exactly the same as fτ , but it has
the same structure and so the proof of convergence is similar). The statement
of [11, Theorem B] only describes smooth convergence along a subsequence
(since they also study the higher dimensional case where bubbling occurs), and
to extend this to show that the limit is unique we use the Lojasiewicz inequality
technique of [19] and [17]. The key estimate is contained in the following
proposition.
implies that
The next step is the main result of this section: The Morse stratification
induced by the gradient flow of fτ is the same as the τ -Harder–Narasimhan
stratification described in Section 8.2.1. First recall the Hitchin–Kobayashi
correspondence from Theorem 8.3.1, and the distance-decreasing result from
[10], which can be re-stated as follows.
Morse theory and stable pairs 157
Lemma 8.3.6 (Hong [10]) Let (A1 , #1 ) and (A2 , #2 ) be two pairs related by
an element g ∈ G C . Then the distance between the G-orbits of (A1 (t), #1 (t))
and (A2 (t), #2 (t)) is non-increasing along the flow.
Recall that the critical sets associated to each stratum are given in Section
8.3.1, and that the critical set associated to the stratum Bδ is denoted ηδ . Define
Sδ ⊂ B to be the subset of pairs that converge to a point in Cδ under the gradient
flow of fτ . The next lemma gives some standard results about the critical sets
of fτ .
Lemma 8.3.7
(i) The critical set ηδ is the minimum of the functional fτ on the stratum Bδ .
(ii) The closure of each G C orbit in Bδ intersects the critical set ηδ .
(iii) There exists ε > 0 (depending on τ ) such that (A, #) ∈ ηδ and (A , # ) ∈
ηδ with δ = δ implies that (A, #) − (A , # ) ≥ ε.
Proof Since these results are analogous to standard results for the Yang–Mills
functional (see for example [1], [6], or [7]), and the proof for holomorphic pairs
is similar, we only sketch the idea of the proof here.
r The first statement follows by noting that the convexity of the norm-square
function · 2 shows that the minimum of fτ on each extension class occurs
at a critical point. This can be checked explicitly for each of the types Ia , Ib ,
II+ , and II− .
r To see the second statement, simply scale the extension class and apply
Theorem 8.3.1 (the Hitchin–Kobayashi correspondence) to the graded object
of the filtration (cf. [7, Theorem 3.10] for the Yang–Mills case).
r The third statement can be checked by noting that (modulo the G-action) the
critical sets are compact, and then explicitly computing the distance between
distinct critical sets.
As a consequence we have
Proposition 8.3.8
(i) Each critical set ηδ has a neighborhood Vδ such that Vδ ∩ Bδ ⊂ Sδ .
(ii) Sδ ∩ Bδ is G C -invariant.
Proof Proposition 8.3.5 implies that there exists a neighborhood Vδ of each
critical set ηδ such that if (A, #) ∈ Vδ then the flow with initial conditions
(A, #) either flows below ηδ , or converges to a critical point close to ηδ . Since
fτ is minimized on each Harder–Narasimhan stratum Bδ by the critical set ηδ ,
the flow is generated by the action of G C , and the strata Bδ are G C -invariant,
then the first alternative cannot occur if (A, #) ∈ Bδ ∩ Vδ . Since the critical
158 R. A. Wentworth and G. Wilkin
sets are a finite distance apart, then (by shrinking Vδ if necessary) the limit
must be contained in ηδ . Therefore Vδ ∩ Bδ ⊂ Sδ , which completes the proof
of (i).
To prove (ii), for each pair (A, #) ∈ Sδ ∩ Bδ , let
Y(A,#) = g ∈ G C : g · (A, #) ∈ Sδ ∩ Bδ .
The aim is to show that Y(A,#) = G C . Firstly we note that since the group
of components of G C is the same as that for the unitary gauge group G, the
flow equations (8.6) are G-equivariant, and the critical sets ηδ are G-invariant,
then it is sufficient to consider the connected component of G C containing the
identity. Therefore the problem reduces to showing that Y(A,#) is open and
closed. Openness follows from the continuity of the group action, the distance-
decreasing result of Lemma 8.3.6, and the result in part (i). Closedness follows
by taking a sequence of points {gk } ⊂ Y(A,#) that converges to some g ∈ G C ,
and observing that the distance-decreasing result of Lemma 8.3.6 implies that
the flow with initial conditions g · (A, #) must converge to a limit close to the
G-orbit of the limit of the flow with initial conditions gk · (A, #) for some large
k. Since the critical sets are G-invariant, and critical sets of different types are a
finite distance apart, then by taking k large enough (so that gk · (A, #) is close
enough to g · (A, #)) we see that the limit of the flow with initial conditions
g · (A, #) must be in ηδ . Therefore Y(A,#) is both open and closed.
Theorem 8.3.9 The Morse stratification by gradient flow is the same as the
Harder–Narasimhan stratification in Definition 8.2.9.
Remark 8.3.10 While we have identified the stable strata of the critical sets
with the Harder-Narasimhan strata, the ordering on the set τ,d coming from the
values of YMH is more complicated. Since this will not affect the calculations,
we continue to use the ordering already defined in Section 2.
We may now reformulate the main result, Theorem 8.1.1. The key idea
is to define a substratification of {Xδ , Xδ }δ∈τ,d by combining Bδ and Aj (δ)
for δ ∈ +τ,d ∩ Iτ,d . In other words, this is simply {Xδ }δ∈τ,d . We call this the
modified Morse stratification.
Morse theory and stable pairs 159
To see (II+ ) and (II− ), we need to compute the solutions to (8.10) and (8.11),
which involves knowing the value of i(μ + iτ · id) on the critical set. Equation
(8.4) shows that
λ1 0
i(μ + iτ · id) =
0 λ2
Proof The first case (when ϕ1 = 0) is easy, since the equations for a ∈
0,1 (L∗1 L2 ) and ϕ2 ∈ 0 (L2 ) become
In the second case (when ϕ1 = 0 is fixed), note (8.12) implies that H(aϕ1 ) =
0, where H denotes the harmonic projection 0,1 (L2 ) → H 0,1 (L2 ). Hence, it
suffices to show that the map
0 (L∗1 L2 ))
D1
/ 0,1 (L∗ L2 ) ⊕ 0 (L2 ) D2
/ 1 (L2 ) (8.18)
1
Remark 8.3.15 The strata for δ ∈ Iτ,d have two components corresponding
to the strata Aj (δ) and Bδ . When there is a possible ambiguity, we will dis-
tinguish these by the notation νI,δ for the negative normal spaces to strata of
type Ia or Ib , and νI I,δ for the negative normal spaces to strata of type II+
or II− .
dimension. In the present case, they are not even linear. In order to carry out
the computations, we appeal to a relative sequence by considering special
subspaces with better behavior.
Definition 8.3.16 For δ ∈ + τ,d ∩ (τ − d/2, τ ], let νI,δ be the negative normal
space to a critical set with # ≡ 0, as in Section 8.3.3. Define
νI,δ = (a, ϕ1 , ϕ2 ) ∈ νI,δ : (a, ϕ1 , ϕ2 ) = 0
νI,δ = (a, ϕ1 , ϕ2 ) ∈ νI,δ : a = 0
Proposition 8.3.17
∗−2(2j (δ)−d+g−1)
δ ∈ + ∗
τ,d ∩ (τ − d/2, τ ] : HG (νI,δ , νI,δ ) HS 1 ×S 1 (ηjA(δ) )
(8.19)
δ∈ +
τ,d ∩ (2τ − d, τ ] : HG∗ (νI,δ
, νI,δ )
∗−2(2j (δ)−d+g−1) d−j (δ)
HS 1 (S M × Jj (δ) (M))
(8.20)
∗−2j (δ)
δ ∈ + ∗
τ,d ∩ (τ −d/2, 2τ −d) : HG (νI,δ , νI,δ ) HS 1 (S j (δ) M ×Jd−j (δ) (M))
(8.21)
∗−2(2j (δ)−d+g−1)
⊕ HS 1 (S d−j (δ) M × Jj (δ) (M))
Proof Fix E = L1 ⊕ L2 . Consider first the case τ > deg L1 = j (δ) > d/2,
and deg L2 = d − j (δ) < d/2. Define the following spaces
ωδ = (A1 , A2 , a, ϕ1 , ϕ2 ) ∈ νI,δ : (a, ϕ2 ) = 0
Zδ− = (A1 , A2 , a, ϕ1 , ϕ2 ) ∈ νI,δ : ϕ1 = 0
Zδ = (A1 , A2 , a, ϕ1 , ϕ2 ) ∈ νI,δ : ϕ1 = 0, (a, ϕ2 ) = 0
Yδ = (A1 , A2 , a, ϕ1 , ϕ2 ) ∈ νI,δ : ϕ1 = 0
Yδ = (A1 , A2 , a, ϕ1 , ϕ2 ) ∈ νI,δ : ϕ1 = 0, (a, ϕ2 ) = 0
Tδ = (A1 , A2 , a, ϕ1 , ϕ2 ) ∈ νI,δ : ϕ1 = 0, (a, ϕ2 ) = 0
Morse theory and stable pairs 163
..
.
··· / H p (νI,δ , ν ) / H p (νI,δ ) / H p (ν ) / ···
G G I,δ
I,δ
o 7 G
ξ oooo
ooo
ooo
p
HG (νI,δ , νI,δ )
OOO
OOOξ
OOO
OO'
... / H p (ν , ωδ ) / H p (ν , ν ) β / H p (ωδ , ν ) / ···
G I,δ G I,δ I,δ G I,δ
..
.
(8.22)
r First, it follows as in the proof of [8, Thm. 2.3] that the pair (νI,δ , νI,δ
) is
A A
homotopic to the Atiyah–Bott pair (Xj (δ) , Xj (δ)−1 ). Hence, (8.19) follows
from [1].
r Consider the pair (νI,δ
, ωδ ). Excision of Zδ gives the isomorphism
HG∗ (νI,δ
, ωδ ) ∼
= HG∗ (νI,δ
\ Zδ , ωδ \ Zδ ) ∼
= HG∗ (Yδ , Yδ ) (8.23)
The space Yδ = Yδ \ Tδ , and Lemma 8.3.12 shows that Yδ is a vector bundle
over Tδ with fibre dimension = deg L1 . Therefore the Thom isomorphism
implies
and therefore
∗−2j (δ)
HG∗ (νI,δ
, ωδ ) = HG∗ (Yδ , Yδ ) = HS 1 (S j (δ) M × Jd−j (δ) (M)) (8.24)
r Consider (ωδ , νI,δ
). By retraction, the pair is homotopic to the intersection
with ϕ1 = 0. It then follows exactly as in [8] (or the argument above) that
)∼
∗−2(2j (δ)−d+g−1) d−j (δ)
HG∗ (ωδ , νI,δ
= HS 1 (S M × Jj (δ) (M)) (8.25)
It then follows as in [8] that ξ , and hence also β, is surjective. This implies
that the lower horizontal exact sequence splits, and (8.21) follows from (8.24)
and (8.25). This completes the proof in this case. The case where deg L1 > τ
is simpler, since ϕ1 ≡ 0 from (8.11). Hence, ωδ = νI,δ , and the proof proceeds
as above.
For all δ ∈ +
τ,d ∩ Iτ,d ,
For all δ ∈ +
τ,d ∩ Iτ,d , δ = τ − d/2,
Eq. (8.28) also holds for δ = τ − d/2, provided d > 4g − 4. In the statements
above, δ1 denotes the predecessor of δ in τ,d .
Remark 8.3.19 Notice that by Corollary 8.3.14 the same argument also proves
(8.27). For d > 4g − 4, we can use Lemma 8.2.17 in the same way to derive
(8.28) for δ = τ − d/2. In this case, by the Thom isomorphism, we have
∗−2(d+2−2g)
HG∗ (Xτ −d/2 , Xτ −d/2 ) HG (Ass ) (8.30)
∼
=
HG (νδ , νδ )
p β
/ H p (ηδ )
G
We first note that the pair (Xδ , Xδ ) is not necessarily invariant under scaling t#,
t → 0, in particular because of the strata in − τ,d (cf. Lemma 8.2.12). However,
166 R. A. Wentworth and G. Wilkin
if we set
,
3δ = Xδ ∪
X 3δ = X
XδA +τ , X 3δ \ pr−1 (Aj (δ) )
δ ≤δ, δ ∈−
τ,d
3δ , X
it follows that HG∗ (Xδ , Xδ ) = HG∗ (X 3δ ). Then for the pair (X
3δ , X
3δ ), pro-
jection to A is a deformation retraction (by scaling the section #), and we
have
3δ ), pr(X
HG∗ (Xδ , Xδ ) = HG∗ (pr(X 3δ )) (8.36)
Next, let
, , ,
Kδ = pr X3(τ −d/2) ∪ Bδ ∪ Bδ ∩ Ak
δ ≤δ , δ ∈−
τ,d δ <τ −d/2 , δ ∈+
τ,d
k>j (δ)
Note that Kδ ⊂ pr(X 3δ ). We claim that it is actually a closed subset of pr(X
3δ ).
Indeed, suppose (Ai , #i ) ∈ X(τ −d/2) , (Ai , #i ) → (A, #) ∈ X 3δ , and suppose
that μ+ (Ai ) > j (δ) for each i. By semicontinuity, it follows that μ+ (A) >
j (δ). On the other had, either A ∈ Kδ or (A, #) ∈ Bδ , τ − d/2 < δ ≤ δ and
δ ∈ +τ,d . But by Lemma 8.2.12, this would imply A ∈ Aj (δ ) ; which is a
contradiction, since j (δ ) ≤ j (δ). It follows that the latter cannot occur, and
hence, Kδ is closed. Similarly,
, ,
3δ ) = pr X
pr(X 3(τ −d/2) ∪ Bδ ∪ Bδ
δ ≤δ , δ ∈−
τ,d δ <τ −d/2 , δ ∈+
τ,d
, ,
∪ Ak ∪ pr(Bδ )
d/2<k≤j (δ) τ −d/2<δ ≤δ , δ ∈+
τ,d
,
= Kδ ∪ Ak
d/2≤k≤j (δ)
f g ∼
=
··· / H p (Xδ1 , X ) / H p (Xδ1 ) / H p (X ) / ···
G δ G G δ
where f and g are induced by the inclusion Xδ1 → Xδ . By Lemma 8.3.20 and
(8.27) (see Remark 8.3.19), it follows that g is surjective and
∗−2j (δ) ∗−2j (δ)
ker g = HG∗ (νI I,δ , νI I,δ ) HG (Bδ ) HS 1 (S j (δ) M × Jd−j (δ) M)
by Thom isomorphism. Chasing through the diagram, it follows that f is also
surjective with the same kernel. We conclude that
∗−2j (δ)
HG∗ (Xδ , Xδ ) HG∗ (Xδ1 , Xδ ) ⊕ HS 1 (S j (δ) M × Jd−j (δ) M) (8.38)
It remains to compute the first factor on the right hand side. To begin, notice
that
, , ,
Ak ∪ Bδ ∪ Bδ
d/2<k<j (δ) τ −d/2<δ <δ , δ ∈−
τ,d τ −d/2<δ <δ , δ ∈+
τ,d
is contained in Xτ −d/2 \ pr−1 (Aj (δ) ) and closed in Xτ −d/2 . This is clear for Ass .
More generally, if (E, #) in this set and # ≡ 0, then μ+ (E) > τ > j (δ), and
elements in the strata of type II− cannot specialize to points in II+ . Again
applying excision, we have
HG∗ (Xδ1 , Xδ ) HG∗ (Yδ , Yδ \ pr−1 (Aj (δ) ))
168 R. A. Wentworth and G. Wilkin
where
,
Yδ = Bss
τ
∪ Bδ
0<δ ≤τ −d/2 , δ ∈+
τ,d
Notice that
τ ,
Bss ∪ Bδ \ Dδ = Bss
j (δ)
We conclude that
HG∗ (Xδ1 , Xδ ) HG∗ (Bss
j (δ)
, Bss
j (δ)
\ pr−1 (Aj (δ) ))
Choose ε > 0 small, and let τ = j (δ) − ε. Then with respect to the τ -
τ
τ
stratification, the right hand side above is HG∗ (Bss ∪ Bετ , Bss ) where ε ∈ −
τ
is the lowest τ -critical set. Since ε < τ − d/2, it follows from Lemma 8.3.20
that the long exact sequence (8.9) splits for this stratum. Hence, we have
∗−2(2j (δ)−d+g−1)
HG∗ (Xδ1 , Xδ ) HG∗ (Bss
τ
∪ Bετ , Bss
τ
) HS 1 (S d−j (δ) M ×Jj (δ) (M))
(8.39)
(notice that jτ (ε) = jτ (δ)). Eqs. (8.38) and (8.39), combined with Propo-
sition 8.3.17, complete the proof. In case δ ∈ Iτ,d , note that by definition
HG∗ (Xδ , Xδ ) HG∗ (Xδ1 , Xδ ). The part of the proof following (8.38) now applies
verbatim to this case.
Proof of Theorem 8.3.18 For δ ∈ + τ,d ∩ [τ − d/2, τ ], or δ = τ − d/2 and
d > 4g − 4, we have proven the result directly (see the discussion following
Theorem 8.3.18 and also Remark 8.3.19). For δ ∈ + τ,d ∩ (τ − d/2, τ ], the
result follows from Proposition 8.3.21 and the five lemma.
Figure 8.2
Proof The first isomorphism is contained in (8.33). To see the second iso-
morphism, note that the results of the last section show that HG∗ (Xδ1 , Xδ ) ∼ =
τ
τ
HG∗ (Bss ∪ Bετ , Bss ), where ε ∈ −
τ is the lowest τ
critical set. Excise all but
a neighborhood of Bετ , and deformation retract # so that # is small. Call
these new sets W and W0 , respectively. Then
τ
τ ∼
HG∗ (Bss ∪ Bετ , Bss ) = HG∗ (W, W0 )
Since # = 0, then we can apply Lemma 8.2.16 to the slices within the
spaces W and W0 , and the resulting spaces are homeomorphic to ωδ and νI,δ
respectively.
170 R. A. Wentworth and G. Wilkin
Figure 8.3
−
The previous lemma together with the surjection ξ : HG∗ (νI,δ
, νI,δ )→
HG∗ (ωδ , νI,δ
)
from (8.22) implies that the map ξg is surjective in the following
commutative diagram (Figure 8.3).
The isomorphism (8.35) together with the results of [1] show that the map
ξg is injective, and so the same argument as before shows that the horizontal
long exact sequence splits.
Proof Since τ is generic, it suffices to prove the result on the level of moduli
spaces, i.e. that the inclusion ı : Mτmax ,d → M τ̃ ,d̃ induces a surjection in
max
cohomology. Consider the determinant map (E, #) → det E. We have the
following diagram
Mτmax ,d
ı /M
(8.40)
τ̃max ,d̃
det det
Jd (M)
j
/ Jd̃ (M)
is surjective onto (det)∗ (H ∗ (Jd (M))). It remains to show that the 2-dimensional
class is in the image of ı ∗ . But since ı is holomorphic and Mτ̃max ,d̃ is projective,
the Kähler class of Mτ̃max ,d̃ restricted to the image generates the cohomology
of the fiber.
Lemma 8.3.24 Suppose δ ∈ τmax ,d , δ < τmax − d/2. Then the inclusion
Xδ → Xδ induces a surjection in G-equivariant rational cohomology. The
same holds for Xτ −d/2 → X . τ̃ −d̃/2
Proof By Lemma 8.3.23, the result holds for the semistable stratum. Fix
δ < τ − d/2, and let δ1 be its predecessor in τmax ,d . By induction, we may
assume the result holds for δ1 . By Lemma 8.3.20 we have the following diagram:
0 / H p (X
δ , X
δ1 ) / H p (X
δ ) / H p (X
δ1 ) /0 (8.41)
G G G
f g h
0 / H p (Xδ , Xδ1 ) / H p (Xδ ) / H p (Xδ1 ) /0
G G G
τ̃ −d̃/2 )
p
HG (X
/ HGp (X ) /0 (8.42)
max τ̃max −d̃/2
h
p
HG (Xτmax −d/2 ) / HGp (Xτ ) / ···
max −d/2
Lemma 8.3.26 Suppose the inclusion Xτ −d/2 → Xτ −d/2 induces a surjection
in G-equivariant rational cohomology for τ = τmax . Then the same is true for
Morse theory and stable pairs 173
all τ ∈ (d/2, d). Moreover, dim HG (Xτ −d/2 , Xτ −d/2 ) is independent of τ for
p
all p.
Proof The sets Xτ −d/2 , Xτ −d/2 remain unchanged for τ in a connected compo-
nent of (d/2, d) \ Cd , where Cd is given in (8.6). Fix τc ∈ Cd , 2τc − d/2 = k ∈
Z, and let τl < τc < τr be in components (d/2, d) \ Cd containing τc in their
closures. Let δ l,r = 2τc − d/2 − τl,r . Note that δ l,r ∈ −τl,r ,d , δ > τl − d/2,
l
Xτr −d/2 = Xτl −d/2 ∪ Bδτll , Xτ r −d/2 = Xτ l −d/2 ∪ Bδτll (8.43)
To see this, we refer to Figure 8.1 and the discussion preceding it. Under
the map τl ,d → τr ,d , δ l → δ r and τl − d/2 → τr − d/2. The claim then
follows if we show that δ r is the predecessor of τr − d/2 in τr ,d , and δ l is
the successor of τl − d/2 in τl ,d (see Figure 8.1). So suppose δ ∈ τr ,d , δ <
τr − d/2. By Remark 8.2.7, we may assume δ ∈ − τr ,d . Write δ + τr = ∈ Z.
Then ≤ 2τr − d/2, which implies ≤ k, and δ ≤ δr . The reasoning is similar
for δl .
Now since the result holds by assumption for τmax , we may assume by
induction that the result holds for τ ≥ τr . Then we have
0 / HGp (Xτr −d/2 , Xτ −d/2 ) / H p (Xτr −d/2 ) / HGp (Xτ −d/2 ) /0
r G r
f g h
··· / HGp (Xτl −d/2 , Xτ −d/2 ) / H p (Xτl −d/2 ) / HGp (Xτ −d/2 ) / ···
l G l
(8.44)
By (8.43) and the proof of Lemma 8.3.24, h is surjective. Hence, the lower long
exact sequence must split. Moreover, g is surjective as well, and ker g = ker h.
As a consequence, f must be an isomorphism. The result now follows by
induction.
Proof of Theorem 8.1.3 This follows from Kirwan surjectivity, but more gen-
erally we prove this on each stratum. Clearly it suffices to prove the result for
174 R. A. Wentworth and G. Wilkin
k = 1. Since the gauge groups for E and E are canonically isomorphic, it suf-
fices by induction to show that if the result holds for the inclusion Xδ → X δ ,
then it also holds for Xδ1 → Xδ1 , where δ1 is the predecessor of δ in τ,d .
By Theorem 8.3.11, the diagram (8.41) holds for all δ. It follows that if g is
surjective, then so is h. This completes the proof.
t 2j (δ)
− Pt S j (δ) M × Jd−j (δ) (M) (8.2)
1−t 2
j t 2(2j (δ)−d+g−1)
Ib (t) = Pt Jj (δ) (M) × Jd−j (δ) (M)
(1 − t ) 2 2
(8.3)
t 2(2j (δ)−d+g−1) d−j (δ)
− Pt S M × Jj (δ) (M)
1−t 2
Morse theory and stable pairs 175
(II+ ) These strata are indexed by integers j = j (δ) = deg # = deg L1 such
that d − N + 1 ≤ j ≤ N − 1, and δ = j − d + τ . The contribution is
t 2j (δ)
II+
j (t) = Pt S j (δ) M × Jd−j (δ) (M) (8.4)
1−t 2
PtG (Bss
τ
) = Pt (BG) − PtG (Xδ , Xδ1 )
δ∈τ,d \{0}
If δ ∈ +
τ,d ∩ (τ − d/2, τ ], then by the Morse–Bott lemma (8.26) and (8.29),
t 2σ (δ) G
PtG (Xδ , Xδ1 ) = P (ηδ )
1 − t2 t
The first term on the right hand side is given by (8.35). For the second term,
we have
6
∗ HG∗ (νI,δ
, νI,δ ) δ ∈ +
τ,d ∩ [2τ − d, τ ]
HG (Xδ1 , Xδ ) =
HG (ωδ , νI,δ ) δ ∈ +
∗
τ,d ∩ (τ − d/2, 2τ − d)
176 R. A. Wentworth and G. Wilkin
and the latter cohomology groups have been computed in (8.20) and (8.25).
This completes the computation.
N−1 ∞
PtG (Bss
j
N
) = Pt (BG) − Id/2
a (t) − Ija (t) − Ib (t)
j =!d/2+1" j =N
d−N−1 N−1
− II−
j (t) − II+
j (t)
j =0 j =d−N+1
!d/2"
1 t 2j − t 2(d+g−1−2j )
a (t) =
Id/2 PtG (Ass ) − Pt (S j M × Jd−j (M))
1−t 2
j =0
1 − t 2
⎧
⎨0 if d odd
− t 2g−2
⎩ Pt (S d/2 M × Jd/2 (M)) if d even
(1 − t 2 )
Remark 8.4.4 It can be verified directly that for d > 4g − 4, the expression
above agrees with (8.1). See the argument of Zagier in [20, pp. 336–7].
Proof of Lemma 8.4.3 Take the special case N = d. Then Mτ,d is a projective
bundle over Jd (M), and so
1 − t 2(d+g−1)
Pt (Mτ,d ) = Pt (Jd (M))
1 − t2
Morse theory and stable pairs 177
d−1 ∞ d−1
Ib (t) − II− II+
j
Pt (Mτ,d ) = Pt (BG) − Id/2
a (t) − Ija (t) − 0 (t) − j (t)
j =!d/2+1" j =d j =1
d−1
t 2j
+ P (S j M × Jd−j (M))
2) t
j =!d/2+1"
(1 − t
d−1
t 2(2j −d+g−1)
+ Pt (S d−j M × Jj (M))
j =!d/2+1"
(1 − t 2)
d−1
t 2j
− Pt (S j M × Jd−j (M))
j =1
(1 − t 2 )
∞
t 2(2j −d+g−1)
= Pt (BG) − Id/2
a (t) − Pt (Jj (M) × Jd−j (M))
j =!d/2+1"
(1 − t 2 )2
d−1
t 2(2j −d+g−1)
+ Pt (S d−j M × Jj (M))
j =!d/2+1"
(1 − t 2)
!d/2"
t 2j
− Pt (S j M × Jd−j (M))
j =1
(1 − t 2 )
Now make the substitution j → d − j in the second to the last sum, using
⎧
⎨d/2 − 1 = !d/2" − 1 if d even
d − !d/2 + 1" =
⎩d/2 − 1/2 = !d/2" if d odd
The result now follows from this, [1, Thm. 7.14], and the fact that Pt (Mτ,d ) is
equal to the j = 0 term in the sum.
178 R. A. Wentworth and G. Wilkin
(8.9)
Morse theory and stable pairs 179
t 2(2N−d+g−1) d−N
+ Pt S M × JN (M)
1 − t2
t 2(2N−d+g−1)
+ Pt (JN (M) × Jd−N (M))
(1 − t 2 )2
t 2(2N−d+g−1) d−N
− Pt S M × JN (M)
1−t 2
t 2(d−2(d−N)+g−1) d−N
+ Pt S M × JN (M)
1−t 2
t 2(d−N) d−N
− Pt S M × JN (M)
1−t 2
t 2N
− Pt S N M × Jd−N (M)
1 − t2
1
= Pt S d−N M ×JN (M) t 4N−2d+2g−2 −t 2d−2N
1 − t2
as required. Using the results of [16] on the cohomology of the symmetric
product, and the fact that Pt (JN (M)) = (1 + t)2g , we see that the same method
as for the proof of [20, (4.1)] gives equation (8.7).
Remark 8.4.7 For τ as above, Theorem 8.4.2 shows that the difference
t 4N−2d+2g−2 d−N
PtG (Bss
N
) − Pt (Mτ,d ) = II−
d−N (t) = Pt S M × JN (M)
1 − t2
comes from only one critical set; the type II critical set corresponding to a
solution of the vortex equations when τ = N . The rest of the terms in (8.9),
corresponding to the difference
+ +
Pt (Mτ +1,d ) − PtG (Bss
N
) = −IN
a (t) + Ib (t) − IId−N (t) − IIN (t)
N
t 2N
=− Pt S N M × Jd−N (M)
1−t 2
come from a number of changes that occur in the structure of the critical
sets as τ increases past N : the term −II+
d−N (t) corresponds to the type II
180 R. A. Wentworth and G. Wilkin
critical point that no longer is a solution to the vortex equations, the term
−II+ +
N (t) corresponds to the new critical point of type II that appears, and the
term −Ia (t) + Ib (t) corresponds to the critical point that changes type from Ib
N N
to Ia .
Therefore we see that the changes in the critical set structure as τ crosses
the critical value N are localized to two regions of B. The first corresponds to
interchanging critical sets of type II− and type II+ . This is the phenomenon
illustrated in Figure 1. The second corresponds to critical sets of type Ia and
II+ that merge to form a single component of type Ib . The terms from the first
change exactly correspond to those in (8.6), i.e.
b (t) − Ia (t) −
and the terms from the second change cancel each other, i.e. IN N
II+
N (t) = 0.
References
[1] M. F. Atiyah and R. Bott. The Yang-Mills equations over Riemann surfaces. Philos.
Trans. Roy. Soc. London Ser. A, 308 (1983), 523–615.
[2] R. Bott. Nondegenerate critical manifolds. Ann. of Math., 60 (1954), no. 2, 248–
261.
[3] S. Bradlow. Special metrics and stability for holomorphic bundles with global
sections. J. Differential Geom. 33 (1991), no. 1, 169–213.
[4] S. Bradlow and G.D. Daskalopoulos. Moduli of stable pairs for holomorphic
bundles over Riemann surfaces. Internat. J. Math. 2 (1991), no. 5, 477–513.
[5] S. Bradlow, G.D. Daskalopoulos, and R.A. Wentworth. Birational equivalences of
vortex moduli. Topology 35 (1996), no. 3, 731–748.
[6] G.D. Daskalopoulos. The topology of the space of stable bundles on a compact
Riemann surface. J. Differential Geom. 36 (1992), no. 3, 699–746.
[7] G.D. Daskalopoulos and R.A. Wentworth. Convergence properties of the Yang-
Mills flow on Kähler surfaces. J. Reine Angew. Math. 575 (2004), 69–99.
[8] G.D. Daskalopoulos, R.A. Wentworth, J. Weitsman, and G. Wilkin. Morse theory
and hyperkähler Kirwan surjectivity for Higgs bundles. J. Differential Geom. 87
(2011), no. 1, 81–116.
[9] S. K. Donaldson. Anti self-dual Yang-Mills connections over complex algebraic
surfaces and stable vector bundles. Proc. London Math. Soc. (3) 50(1) (1985),
1–26.
[10] M.-C. Hong. Heat flow for the Yang-Mills-Higgs field and the Hermitian Yang-
Mills-Higgs metric. Ann. Global Anal. Geom. 20(1) (2001), 23–46.
Morse theory and stable pairs 181
[11] M.-C. Hong and G. Tian. Asymptotical behaviour of the Yang-Mills flow and
singular Yang-Mills connections. Math. Ann. 330(3) (2004), 441–472.
[12] A. King and P. Newstead. Moduli of Brill-Noether pairs on algebraic curves.
Internat. J. Math. 6 (1995), no. 5, 733–748.
[13] F. C. Kirwan. “Cohomology of quotients in symplectic and algebraic geometry”,
volume 31 of Mathematical Notes. Princeton University Press, Princeton, NJ,
1984.
[14] S. Kobayashi. “Differential geometry of complex vector bundles”, volume 15 of
Publications of the Mathematical Society of Japan. Princeton University Press,
Princeton, NJ, 1987. , Kano Memorial Lectures, 5.
[15] J. Le Potier. Systèmes cohérents et structures de niveau. Astérisque No. 214 (1993),
143 pp.
[16] I. G. Macdonald. Symmetric products of an algebraic curve. Topology 1 (1962),
319–343.
[17] J. Råde. On the Yang-Mills heat equation in two and three dimensions. J. Reine
Angew. Math. 431 (1992), 123–163.
[18] N. Raghavendra and P. Vishwanath. Moduli of pairs and generalized theta divisors.
Tohoku Math. J. (2) 46 (1994), no. 3, 321–340.
[19] L. Simon. Asymptotics for a class of nonlinear evolution equations, with applica-
tions to geometric problems. Ann. of Math. (2) 118(3) (1983), 525–571.
[20] M. Thaddeus. Stable pairs, linear systems and the Verlinde formula. Invent. Math.
117, no. 2 (1994), 317–353.
[21] G. Wilkin. Morse theory for the space of Higgs bundles. Comm. Anal.Geom. 16,
no. 2 (2008), 283–332.
Authors’ addresses:
Richard A. Wentworth
Department of Mathematics,
University of Maryland,
College Park, MD 20742, USA
[email protected]
Graeme Wilkin
Department of Mathematics
University of Colorado
Boulder, CO 80309
USA
[email protected]
9
Manifolds with k-positive Ricci curvature
jon wolfson
9.1 Introduction
Let (M, g) be an n-dimensional Riemannian manifold. We say M has k-positive
Ricci curvature if at each point p ∈ M the sum of the k smallest eigenvalues of
the Ricci curvature at p is positive. We say that the k-positive Ricci curvature
is bounded below by α if the sum of the k smallest eigenvalues is greater than
α. Note that n-positive Ricci curvature is equivalent to positive scalar curvature
and one-positive Ricci curvature is equivalent to positive Ricci curvature. We
first describe some basic connect sum and surgery results for k-positive Ricci
curvature that are direct generalizations of the well known results for positive
scalar curvature (n-positive Ricci curvature). Using these results we construct
examples that motivate questions and conjectures in the cases of 2-positive and
(n − 1)-positive Ricci curvature. In particular:
182
Manifolds with k-positive Ricci curvature 183
space theory, like in the Bonnet–Myers argument, as well as some notions from
geometric group theory. See [R-W] for full details and Section 4 of this chap-
ter for an overview. However, the question: “Does a manifold with 2-positive
Ricci curvature bounded below by α satisfy an upper bound on its fill radius?”
remains open.
It is clear that k-positive Ricci curvature implies (k + 1)-positive Ricci
curvature, so results on n-manifolds with n-positive Ricci curvature hold for n-
manifolds with k-positive Ricci curvature, any k, 0 < k < n. As a result of the
surgery theorem we pose some interesting questions on the difference between
n-positive Ricci curvature and (n − 1)-positive Ricci curvature.
The author was partially supported by NSF grant DMS-0604759.
change the metric in D preserving that the Ricci curvature is k-positive such
that the metric agrees with the original metric near ∂D and such that near p
the metric is the standard metric on S (n−1) (ε) × R, for any ε sufficiently small.
It follows from this that 1-handles can be added and connected sums taken
preserving that the Ricci curvature is k-positive, for any 2 ≤ k ≤ n.
The method of Gromov–Lawson proceeds as follows: We consider the Rie-
mannian product D × R with coordinates (x, t), where x are normal coordinates
on D. We define a hypersurface N ⊂ D × R by the relation
N = {(x, t) : (x, t) ∈ γ }
R ij ij = Rij ij − λi λj for i = j
R ij lj = Rij lj for i = l.
we have:
R ij lj = RijDlj , for i, j, l = 2, . . . , n,
R 1j ij = R D∂ j ij cos θ, for i, j = 2, . . . , n,
∂r
R i1j 1 = RiD∂ j ∂ cos2 θ, for i, j = 2, . . . , n,
∂r ∂r
where R D is the curvature tensor of the metric on D. It follows that the Ricci
curvature of N at (x, t) with respect to the frame {e1 , . . . , en } is given by:
n
Ric11 = RicD ∂
, ∂
∂r ∂r
cos2 θ + κ λj (9.2.1)
j =2
Ric1j = RicD ∂
∂r
, ej cos θ, for j = 2, . . . , n,
Ricij = RicD
ij −RiD∂ j ∂ sin θ, 2
for i, j = 2, . . . , n, i = j,
∂r ∂r
⎛ ⎞
n
Ricii = RicD
ii −Ri ∂ i ∂ sin θ + λi
D 2 ⎝ λj + κ ⎠ , for i = 2, . . . , n,
∂r ∂r
j =2,j =i
Choose r0 > 0 small. From (9.2.2) it follows that the the Ricci curvature of the
hypersurface N has (n − 1) positive eigenvalues and that each of these eigenval-
ues is larger in absolute value than the one remaining eigenvalue (corresponding
to the direction of e1 ). Therefore the Ricci curvature remains k-positive and, in
fact, becomes 2-positive. Consider the point (r0 , t0 ) ∈ γ0 . Bend γ0 , beginning
at this point, with the curvature function κ(s) similar to the one used in [G-L1]:
186 J. Wolfson
κ
sin θ0
2r0
s
1
2 r0
Using (9.2.1) (and r0 sufficiently small) this implies that along the bend the
Ricci curvature of N remains 2-positive. During this bending process the curve
will not cross the line r = r0 /2 since the length of the bend is r0 /2 and it
begins at height r0 . The total amount of bending is:
sin θ0
θ = κds ,
4
independent of r0 .
Continue the curve with a new straight line segment γ1 at an angle θ1 =
θ0 + θ . Repeat the above procedure now using θ1 where previously we used
θ0 . The total bending of this procedure will then be sin4θ1 > sin4θ0 . By repeating
this procedure a finite number of times (depending on sin θ0 ) we can achieve a
total bend of π/2. This completes the proof of the connect sum result.
For the general case of surgeries we again explain the modification of the
Gromov-Lawson argument. Let S p ⊂ M be an embedded sphere with trivial
normal bundle B of dimension q ≥ 3. Identify B with S p × Rq . Define r :
S p × Rq → R+ by r(y, x) = ||x||, and set S p × D q (ρ) = {(y, x) : r(y, x) ≤
ρ}. Choose r̄ > 0 such that the normal exponential map exp : B → M is an
embedding on S p × D q (r̄) ⊂ B. Lift the metric on M to S p × D q (r̄) by the
exponential map. Then r is the distance function to S p × {0} in S p × D q (r̄),
and curves of the form {y} × , where is a ray in D q (r̄) emanating from the
origin, are geodesics in S p × D q (r̄).
We now consider hypersurfaces in the Riemannian product (S p × D q (r̄)) ×
R of the form:
where γ is, as above, a curve in the (r, t)-plane. As in the connect sum case
we must choose γ so that the metric on N has k-positive Ricci curvature at all
points. We first remark that γ = N ∩ ({y} × × R) is a principal curve on N
and the associated principal curvature at a point corresponding to (r, t) ∈ γ is
exactly the curvature κ of γ at that point.
Now fix a point q ∈ γ ⊂ N corresponding to a point (r, t) ∈ γ . Let
{e1 , . . . , en } be an orthonormal basis of Tq (N ) such that e1 is the tangent
vector to γ , and e2 , . . . , en are principal vectors for the second fundamen-
tal form of N . If the metric on S p × D q (r̄) at q is the product metric
then the vectors e2 , . . . , eq can be chosen to be tangent to {y} × D q (r̄) and
the vectors eq+1 , . . . , en to be tangent to S p × {x}. The principal curvatures
λ2 , . . . , λq are then of the form (− 1r + O(r)) sin θ , where θ is the angle
between the normal to the hypersurface and the t-axis and the principal cur-
vatures λq+1 , . . . , λn are of the form O(1) sin θ (i.e., are independent of r).
In the general case such a simple description is not possible. However if r is
small the q − 1 largest principal curvatures are of the form (− rc + O(1)) sin θ ,
where c is a positive constant that can be bounded away from zero and the
remaining principal curvatures are of this form O(1) sin θ . (Since the prod-
uct of the principal curvatures grows like r q−1 1
.) We will denote the q − 1
largest principal curvatures by λ2 , . . . , λq corresponding to the directions
e2 , . . . , eq and the remainder by λq+1 , . . . , λn corresponding to the directions
eq+1 , . . . , en .
The Gauss curvature equation relates the curvature tensor Rij lm of N with
the curvature tensor R ij lm of (S p × D q (r̄)) × R. In particular, with respect to
this basis, at q:
R ij ij = Rij ij − λi λj , for i = j,
R ij lj = Rij lj , for i = l,
R ij lj = RijS lj×D ,
p q
for i, j, l = 2, . . . , n,
R 1j ij = R S∂ j×D
p q
ij
cos θ, for i, j = 2, . . . , n,
∂rp
R i1j 1 = RiS ∂ ×D
q
j ∂
cos2 θ, for i, j = 2, . . . , n,
∂r ∂r
188 J. Wolfson
(Note that since λq+1 , . . . , λn are bounded for r small, we cannot conclude any
more than Ricii = O( 1r ) for i = q + 1, . . . , n.) Therefore, provided q ≥ 3, for
r sufficiently small the Ricci curvature has q − 1 positive eigenvalues that
strongly dominate all other eigenvalues. In particular, provided k > n − q + 1
the Ricci curvature remains k-positive. We can then bend γ to a line parallel to
the t-axis and preserve k-positive Ricci curvature as above.
This construction of N determines a “tube” of k-positive Ricci curvature
with two boundary components. The initial boundary component has a collar
neighborhood isometric to a tubular neighborhood of S p in M. The final bound-
ary component has a collar neighborhood isometric to ∂(S p × D q (ε)) × R =
S p × S q−1 (ε) × R for the product metric with the R factor. This allows us to
Manifolds with k-positive Ricci curvature 189
Proof Consider the metric gf (t) + dt 2 on X × [0, a], where f (t) is a C 2 ([0, a])
function that is monotonically increasing from 0 to 1. We will explic-
itly determine f (t) below. We compute the curvature of gf (t) + dt 2 at the
point (x0 , t0 ) ∈ X × [0, a]. Let e0 be the unit normal along the hypersur-
face X × {t0 } pointing in the direction ∂t∂ . Let {e1 , . . . , en } be an orthonor-
mal frame along X × {t0 } near (x0 , t0 ). Then {e0 , . . . , en } is an orthonormal
190 J. Wolfson
frame along X × [0, 1] near (x0 , t0 ). Let {ω0 , . . . , ωn } be the dual coframe.
For notational convenience we use the index ranges α, β, γ = 0, 1, . . . , n and
i, j, k = 1, 2, . . . , n. The connection one-form for the coframe {ω0 , . . . , ωn }
is {ωαβ }. The one-forms ωij depend only on the metric gf (t) . However the
one-forms ω0i depend on f (t) so at (x0 , t0 ) we have:
ω0i = O(|f |), (9.2.5)
2
dω0i = O(|f |) + O(|f |) + O(|f | ).
The curvature two-form αβ on X × [0, a] is determined by the structure
equation:
αβ = dωαβ − ωαγ ∧ ωγβ . (9.2.6)
γ
admit metrics with 2-positive Ricci curvature. From this we see that compact
manifolds with 2-positive Ricci curvature can have large fundamental groups.
In fact the fundamental groups of the examples (9.2.9) are virtually free. This
implies that the manifolds (9.2.9), when l ≥ 2, do not admit metrics of non-
negative Ricci curvature since the universal covers of (9.2.9) have infinitely
many ends. In contrast the universal cover of a compact manifold with non-
negative Ricci curvature splits isometrically as a product N × Rp , where N is a
compact manifold, and therefore has one or two ends. Many other topologically
distinct examples of compact manifolds that admit metrics of 2-positive Ricci
curvature can be constructed by taking the connect sum of manifolds of positive
Ricci curvature with the manifolds of (9.2.9).
Consider the round metric on the sphere of radius r, Srn−2 , and the hyperbolic
metric on the Riemann surface g of genus g ≥ 2. When r is sufficiently small
the manifolds Srn−2 × g admit metrics of 3-positive Ricci curvature but not
of 2-positive Ricci curvature. Examples of this type indicate that 3-positive
Ricci curvature when n ≥ 4 imposes much weaker restrictions on π1 (M) than
2-positive Ricci curvature.
The surgery results of Theorem 9.2.1 do not allow any surgeries preserving
2-positive Ricci curvature (except connect sum). When k > 2 and n > 3, q =
n − 1-surgeries preserve the curvature condition. This suggests a difference
between k = 2 and k > 2 in the rigidity of the fundamental group for manifolds
with k-positive Ricci curvature.
S 2 × T 2 , where S 2 is the round 2-sphere and T 2 is the flat 2-torus, has 2-non-
negative Ricci curvature and fundamental group isomorphic to Z ⊕ Z.
The surgery statement of Theorem 9.2.1 shows that if k = n − 1 then surgery
is possible provided q ≥ 3. This is the same condition as in the positive scalar
curvature case. The surgery result is used by Gromov-Lawson [G-L1] to prove
that if n ≥ 5 then every compact simply-connected non-spin n-manifold carries
a metric of positive scalar curvature and by Stolz [Sz] to prove that for n ≥ 5
every compact simply-connected spin n-manifold with vanishing α invariant
carries a metric of positive scalar curvature. Gromov-Lawson prove their result
for non-spin manifolds using oriented bordism. Stolz proves his result for spin
manifolds using spin bordism. Since n − 1 positive Ricci curvature implies
positive scalar curvature any necessary condition for positive scalar curvature
is a necessary condition for n − 1 positive Ricci curvature. In light of the
surgery result for n − 1 positive Ricci curvature exactly the same arguments
apply to prove:
To prove the theorem, all that needs to be checked is that the generators of
oriented bordism described in [G-L1] and the HP2 bundles used in [Sz] admit
metrics with n − 1 positive Ricci curvature. In particular for compact simply-
connected n-manifold with n ≥ 5 there is no distinction between positive scalar
curvature and n − 1 positive Ricci curvature.
Question 1 Is there a compact n-manifold, n ≥ 5, that admits a metric of
positive scalar curvature but does not admit a metric with n − 1 positive Ricci
curvature?
where d (−, −) denotes the distance on in the induced metric. Thus the
fill radius of γ can be bounded above by an upper bound on the diameter
of in the induced metric. For arbitrary spanning γ such a bound is, of
course, impossible. However in [G-L2] and [S-Y2], is taken to be an area
minimizer among discs spanning γ (i.e., a solution of the Plateau problem). In
particular, is a stable, minimal immersion. After perturbing γ inward along
the minimal surface is strictly stable for normal variations vanishing on the
boundary. In [S-Y1] the second variation of area for a minimal surface in a
three manifold M is given. Let f ν be a compactly supported normal variation,
194 J. Wolfson
where f ∈ C0∞ () and ν is a unit normal. Then the second variation formula
can be written:
d 2 || 1
2
= |∇f |2 + f 2 K − S − ||A||2 da. (9.3.1)
dt |t=0 2
L(f ) = f + f (K − S) (9.3.2)
I (ψ, ψ) =
dk d 2k 2
− ψ +k −1 ψ +ψ(K +k −1 k +k −1 2 ) ψ + 2 (∇k · ν)2 ψ 2 kds
0 ds ds k
(9.3.3)
Then,
2
I (ψ, ψ) = L0 (ψ)ψ − (∇k · ν)2 ψ 2 kds (9.3.5)
0 k2
g + k −1 g k + g(k −1 L(k) + S + k −1 k ) ≤ 0.
g −1 g + g −1 k −1 g k + S + k −1 k ≤ 0, (9.3.8)
on [0, ]. Let ϕ be any smooth function on [0, ] vanishing at the endpoints and
multiply (9.3.8) by ϕ 2 to give:
(g −1 g ϕ 2 + g −1 k −1 g k ϕ 2 + k −1 k ϕ 2 + Sϕ 2 )ds ≤ 0
0
where the Si are spherical space forms (this uses the solution of the Poincare
conjecture) and the Ki are K(π, 1) manifolds (a K(π, 1) manifold is a closed
manifold with contractible universal cover and fundamental group isomorphic
to a group π .) Using the fill radius bound it can be shown that if, in addition, M
has positive scalar curvature then no K(π, 1) summands occur in this direct sum
decomposition (see [G-L2 ] for a proof). In particular, this implies that if M is a
closed three manifold with positive scalar curvature then the fundamental group
of M is virtually free. This verifies Conjecture 1 in the three dimensional case
since 2-positive Ricci curvature implies 3-positive Ricci curvature (positive
scalar curvature).
Manifolds with k-positive Ricci curvature 197
Theorem 9.3.2 Let (M, g) be a closed n-manifold with fill radius bounded
above by β. Then there is a distance decreasing map φ : M → " onto a metric
graph, such that, for each p ∈ ",
Proof The theorem follows from the proof of Corollary 9.11 in [G-L2]. Also
see [G] Appendix 1.
The theorem implies that closed n-manifolds with fill radius bounded above
and large diameter are “long” and “thin”, exactly like compact 3-manifolds
with positive scalar curvature and large diameter.
We have already observed that in three dimensions 2-positive Ricci curvature
implies 3-positive Ricci curvature (positive scalar curvature). Therefore closed
three manifolds with 2-positive Ricci curvature bounded away from zero by α
satisfy a fill radius bound depending on α. This partly motivates the following
conjecture:
There is another “motivation” for this conjecture: Let (, ∂) be a stable
minimally immersed surface with boundary in a Riemannian four-manifold
(M, g). Using an averaging technique, the Gauss equation and an appropriate
choice of variational vector field, the second variation formula for area and
stability can be used to show that for any smooth function of compact support
f ∈ C0∞ () we have the inequalities:
1
|∇f |2 + f 2 K − Kν − (Ric11 + Ric22 ) eσ da ≥ 0 (9.3.12)
2
and
1
|∇f |2 + f 2 K + Kν − (Ric11 + Ric22 ) e−σ da ≥ 0. (9.3.13)
2
Here K is the Gauss curvature on (in the induced metric), Kν is the curvature
of the normal bundle, Ric is the Ricci curvature on M, {e1 , e2 } is an orthonormal
frame on and σ is a function on that satisfies σ = Kν . Suppose that
(M, g) has 2-positive Ricci curvature bounded below by α > 0. Then,
Ric11 + Ric22 ≥ α.
198 J. Wolfson
In the case that the normal bundle is flat or, more generally, that the function σ
has suitably small oscillation, equations (9.3.12) and (9.3.13) imply:
1
|∇f |2 + f 2 K − (Ric11 + Ric22 ) da ≥ 0. (9.3.14)
2
Then (9.3.14) can be used in the above argument from [S-Y2] to prove a
diameter bound on in the induced metric. This implies a fill radius bound
and hence Conjecture 2. However, it can be shown that this line of reasoning
does not, in general, hold, the obstruction being the normal curvature. This
does not mean that a stable minimally immersed surface in a Riemannian
four-manifold (M, g) with 2-positive Ricci curvature bounded below does not
satisfy a diameter bound. Rather that the above reasoning does not apply. It
remains an interesting, if unexploited, fact that the 2-positive Ricci curvature
occurs in an averaged version of the second variation formula.
In the next section we will show that Conjecture 2 implies Conjecture 1 (in
the four dimensional case).
lie in the same end of M̃. The geodesic line γ̃ consists of two geodesic rays
γ̃1 and γ̃2 beginning at x. For i = 1, 2, choose a point pi ∈ M̃ \ BR (x) along
γ̃i and denote the segment of γ̃i from x to pi by τi . Since p1 and p2 lie in the
same end there is a curve β ⊂ M̃ \ BR (x) joining p1 and p2 . Denote the closed
curve τ1 ∪ β ∪ τ2 by η. Since M̃ is simply connected η is null homotopic and
has fill radius greater than R2 . For sufficiently large R this contradicts the fill
radius bound.
To illustrate the use of Proposition 9.4.4 we prove the following special case
of Theorem 9.4.1.
References
[E] Epstein, D., Ends, Topology of 3-manifolds edited by, M.K. Fort, Jr., Prentice-
Hall, 1962, 110–117.
[Ga] Gajer, J., Riemannian metrics of positive scalar curvature on compact manifolds
with boundary, Ann. Glob. Anal. Geom. 5 (1987) 179–191.
[G] Gromov, M, Filling Riemannian manifolds, J. Diff. Geom. 18 (1983) 1–147.
[G-L1] Gromov, M, and Lawson, H., The classification of simply connected manifolds
of positive scalar curvature, Ann of Math. 111 (1980) 423–434.
[G-L2] Gromov, M, and Lawson, H., Positive scalar curvature and the Dirac operator
on complete Riemannian manifolds, Publ. Math de IHES, 58 (1983) 83–196.
[Ma] Massey, W., Algebraic Topology: An Introduction, GTM 56, Springer-Verlag,
New York, 1984.
[R-W] Ramachandran, M. and Wolfson, J., Fill Radius and the Fundamental Group,
Journal of Topology and Analysis 2 (2010) 99–107.
Manifolds with k-positive Ricci curvature 201
[R-S] Rosenberg, J., and Stolz, S., Metrics of positive scalar curvature and connections
with surgery, in Surveys on Surgery Theory, Volume 2, edited by S. Cappell,
et al., Annals of Math Studies 149, Princeton Univ. Press, Princeton, 2001.
[S-Y1] Schoen, R., and Yau, S. T., On the structure of manifolds of positive scalar
curvature, Manu. Math. 28 (1979) 159–183.
[S-Y2] Schoen, R., and Yau, S. T., The existence of a black hole due to condensation
of matter, Comm. Math. Phys. 90 (1983) 575–579.
[Sz] Stolz, S., Simply connected manifolds of positive scalar curvature, Ann. of
Math. 136 (1992), 511–540.
[St] Stallings, J., On torsion-free groups with infinitely many ends, Ann. of Math.
88 (1968), 312–334.
Author’s address:
Department of Mathematics,
Michigan State University,
East Lansing, MI 48824
USA
[email protected]